pubs.acs.org/jmc Article N‑Acylamino Saccharin as an Emerging Cysteine-Directed Covalent Warhead and Its Application in the Identification of Novel FBPase Inhibitors toward Glucose Reduction Wuqiang Wen,∥ Hongxuan Cao,∥ Yixiang Xu,∥ Yanliang Ren,* Li Rao, Xubo Shao, Han Chen, Lixia Wu, Jiaqi Liu, Chen Su, Chao Peng, Yunyuan Huang,* and Jian Wan* Downloaded via IFECT- DO SUL-RIO-GRANDENSE on October 2, 2023 at 21:45:41 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles. Cite This: J. Med. Chem. 2022, 65, 9126−9143 ACCESS Metrics & More Read Online Article Recommendations sı Supporting Information * ABSTRACT: With a resurgence of covalent drugs, there is an urgent need for the identification of new moieties capable of cysteine bond formation. Herein, we report on the N-acylamino saccharin moieties capable of novel covalent reactions with cysteine. Their utility as alternative electrophilic warheads was demonstrated through the covalent modification of fructose-1,6bisphosphatase (FBPase), a promising target associated with cancer and type 2 diabetes. The cocrystal structure of title compound W8 bound with FBPase unexpectedly revealed that the N-acylamino saccharin moiety worked as an electrophile warhead that covalently modified the noncatalytic C128 site in FBPase while releasing saccharin, suggesting a previously undiscovered covalent reaction mechanism of saccharin derivatives with cysteine. Treatment of title compound W8 displayed potent inhibition of glucose production in vitro and in vivo. This newly discovered reactive warhead supplements the current repertoire of cysteine covalent modifiers while avoiding some of the limitations generally associated with established moieties. ■ INTRODUCTION With the development of modern medicine, targeted covalent inhibitors (TCIs) have become greatly successful therapies for a broad array of human diseases such as non-small-cell lung carcinoma, mantle cell lymphoma, and type II diabetes.1−3 Currently, approximately one-third of targeted enzyme drugs approved by the Food and Drug Administration (FDA) are covalent inhibitors.4,5 Due to the increased strength and often irreversible nature of the covalent bonds formed between a TCI and target, the use of covalent inhibitors offers the potential for increased potency and prolonged pharmacodynamics effects, compared to traditional noncovalent inhibitors.6,7 In addition, targeting disease proteins and pathways with covalent inhibitors has become a feasible option for overcoming drug resistance and protein mutation and improving the protein isoform selectivity and toxicity associated with noncovalent inhibitors.4,8−10 The TCIs bind to target proteins in two distinct necessary steps: the first step involves the reversible binding of a highaffinity ligand to its biological target and then an electrophilic “warhead” on the ligand binds at the appropriate position to form a covalent bond with a nucleophilic residue on the protein.9,11 A number of electrophilic warheads have been explored to react with nucleophile residues, including cysteine, lysine, or tyrosine;12,13 however, cysteine’s thiol is endowed with enhanced reactivity, and the paucity of cysteine in the © 2022 American Chemical Society proteome coupled with the fact that closely related proteins do not necessarily share a given cysteine residue enables a level of unprecedented rational target selectivity, making cysteine the most favored target.14,15 Warhead selection typically starts with an estimation of the reactivity required to target the desired amino acid, as the reactivity profiles of the covalent reactive group (warhead) affect the target specificity of TCIs.16 Nevertheless, it is still challenging to strike the right balance between reactivity and selectivity. The most recently reported warheads undergo both Michael-type and non-Michael-type nucleophilic addition, addition−elimination reaction, nucleophilic substitution, and oxidation.12,17 Targeting noncatalytic cysteine residues with acrylamides and other α,β-unsaturated carbonyl compounds is the “classical” strategy of TCI development.15,18 A recent analysis of cysteine-targeted covalent inhibitors revealed that nearly 70% of the published compounds carried Michael acceptor-type warheads, and acrylamide is the preponderant functional group12 due to its Received: March 2, 2022 Published: July 5, 2022 9126 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article Figure 1. (A) Model reaction of W1 with thiol reagent 1,4-benzenedithiol. Reaction conditions: tetrahydrofuran (THF), N2, rt. (B) Model reaction of W1 with glutathione (GSH, 1.5 equiv) in pH 7.4 D2O/d6-acetone (1/5) at 37 °C. (C) Proposed reaction mechanism of W1 with cysteine. In this study, we report our efforts that sweetener saccharin derivatives can be used to covalently modify the thiols or cysteine(s) of proteins through a previously unreported reaction mechanism. Most particular interest is that this newly discovered warhead not only covalently inhibits target protein but also releases saccharin, which is not metabolized and considered safe by the FDA. Saccharin derivatives have shown inhibitory activity against various different subtypes of carbonic anhydrases in humans by coordinating with their zinc ions;40,41 however, there are many zinc-dependent enzymes in organisms,42 which makes it difficult for these inhibitors to act effectively on their intended target. The covalent reaction of saccharin derivatives that we newly report herein provides a new strategy for achieving the purpose of the precise release of saccharin inhibitors through reaction with glutathione (GSH) or cysteine in the future. The reactivity of this novel warhead was assessed on the basis of GSH reactivity through a quantitative 1H NMR (qNMR) method43 and quantum mechanical (QM) calculations. Promising scaffolds were investigated further in the identification of the inhibitory ability of fructose-1,6-bisphosphatase (FBPase), a key ratecontrolling enzyme in gluconeogenesis44,45 and a promising target associated with cancer and type 2 diabetes.46,47 Our previous studies have identified a new covalent allosteric site (C128) of FBPase,48,49 which provides a promising way for the design of covalent allosteric drugs for glucose reduction, allowing incorporation of an electrophilic group at the appropriate position. Therefore, we sought to design a novel covalent inhibitor of FBPase with an optimal warhead. The covalent binding mode was identified by combining sitedirected mutagenesis, protein liquid chromatography with mass spectrometry (LC−MS), and cocrystal structure analysis. Notably, compounds W8 and W8k exhibited high selectivity against FBPase and W8 effectively reduced blood glucose in an Institute of Cancer Research (ICR) mice model and dosedependent inhibition of glucose production in a primary mouse hepatocyte model. ease of synthesis; acceptable reactivity window with cysteine over other amino acids; and absorption, distribution, metabolism, excretion, toxicity (ADMET) compatibility.19 However, the majority of drugs featuring a covalent binding modality were discovered serendipitously and were only retrospectively identified as covalent inhibitors therefore, only a few drugs that contain the abovementioned “classical” electrophiles have entered clinical practice.5 Recent cytochemical proteomic studies have shown that certain types of acrylamide-based kinase inhibitors induce the expression of offtarget protein markers in the submicromolar concentration range.20 Therefore, there remains a significant need for additional cysteine reactive warheads with tunable properties and reactivity profiles.6 Similarly, many issues with other types of established warheads remain to be resolved. For example, with respect to nucleophilic substitution-based warheads, no studies on sulfonyl fluorides,21,22 nucleophilic aromatic substitution (SNAr) electrophiles,23−25 or activated esters26,27 investigated the toxic potential of the leaving group.17 In addition, problems may arise when the target residue is poorly reactive, difficult to access, or incompatible with the spatial and geometric requirements of these electrophilic warhead groups. This prompted us to search for as-yet unexplored electrophiles to increase the electrophilic warhead options available for TCI design. Saccharin, is an orally effective, noncalorie artificial sweetener (NAS).28 Since saccharin is not metabolized and is considered safe by the FDA,29 it has been used as a substructure in a number of bioactive compounds, including carbonic anhydrase (CA) inhibitors,30−33 leukocyte elastase inhibitors,34 and neutrophil elastase inhibitors.29 Furthermore, compounds containing saccharin fragments are used in the clinic for antidepressant35 or stroke36,37 therapy. Notably, saccharin has recently been reported to be a warhead that can covalently bind to serine through a nucleophilic addition reaction;38,39 however, this type of covalent reaction results in the destruction of the saccharin structure and diminished safety. 9127 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry ■ pubs.acs.org/jmc Article Scheme 1. Synthesis of Compounds W2 and W3a RESULTS AND DISCUSSION Electrophilic Reactivity of W1. Saccharin-based inhibitors have been described as irreversibly acylated compounds of rhomboids (intramembrane serine proteases), and saccharin has been regarded as a warhead that can covalently bind to serine through a nucleophilic addition reaction.37 Specifically, a reaction mechanism in which the serine in the active site of a protease attacks the carbonyl group of saccharin, leading to opening of the ring in the saccharin-based inhibitor and the formation of an acyl-enzyme, has been proposed. Nevertheless, the application of this type of saccharin-based inhibitor is very limited in drug discovery due to its low reactivity of the endocyclic carbonyl group. In this study, we found that the reaction of Nbenzoylsaccharin (W1) with 1,4-benzenedithiol affords saccharin and S-(4-mercaptophenyl) benzothioate (adduct 1, Figures 1A and S1), indicating that W1 may be a potential thiol electrophilic reagent with a novel reaction mechanism that has never been reported. Interestingly, this reaction releases saccharin, a sweetener that is not metabolized and considered safe by the FDA, as well as an inhibitor of CA linked to many diseases such as edema, glaucoma, epilepsy, and cancer.31 Therefore, W1 may be a potential novel covalent warhead covalently binding to the thiol of cysteine. GSH is a tripeptide consisting of glutamic acid, cysteine, and glycine, and it contains a free thiol moiety that acts as a reactive nucleophilic site and has been typically applied to the identification of electrophilic reagents in vitro.23,49,50 On the basis of the reaction (Figure 1A) of W1 and 1,4benzenedithiol, we deduced that it is feasible to use W1 as a novel electrophile to react with GSH, and their probable products (Figure 1B) are Adduct 2 and saccharin. To determine the feasibility of this reaction, the solution of W1 in acetone-d6 was treated with a solution containing 1.5 mol equiv of GSH in pH 7.4 buffered D2O. As expected, the saccharin was isolated from the products. Furthermore, nuclear magnetic resonance (NMR) spectroscopy (Figure S2B) was performed to trace Adduct 2 of this reaction. It is shown in Figure S2B that the peaks corresponding to the methylene (a) of GSH and the phenyl (b) of W1 decrease gradually, while two newly formed peaks (c and d), which are related to the products, increase gradually, suggesting the presence of Adduct 2 proposed in Figure 1B. On the basis of the aforementioned experimental data, we proposed the possible reaction mechanism (Figure 1C) of W1 reacting with a thiol. This reaction starts with a nucleophilic attack on the exocyclic carbonyl group (blue), leading to a tetrahedral alkoxide transition state (Figure S3), which has been predicted at the ωb97xd/6-31+g(d)-SMD level of theory using the Gaussian 09 software package.51 The transition energy barrier (ΔE1) of step 1 is 65.87 kcal/mol. The alkoxide negative charge acquired greater stability after being transferred to the leaving group (saccharin), and the elimination of the leaving group (saccharin) in step 2 allowed the carbonyl reformation of a new acyl compound with cysteine. Our calculated energy barrier (ΔE1) for step 2 was 75.8 kcal/mol. Chemistry. Compound 2 was synthesized by a known methodology.52 Then, nucleophilic substitution of 2 with benzoyl chloride in dichloromethane (DCM) (in the presence of Et3N) at room temperature (rt) yielded W2 (Scheme 1). Compound W3 was prepared using the same reaction route as that for W2. W4−W6 were synthesized using saccharin (1) as a Reagents and conditions: (a) THF, LiAlH4, rt, 1 h; (b) benzoyl chloride, DCM, rt, Et3N, 1.5 h. the starting material with the same reaction route as that for compound W3 (Scheme 2). W7 was prepared as previously described53 (Scheme 2). The formation of compound 8 in the reaction of compound 7 with triphosgene was followed by a reaction between this intermediate compound and saccharin to afford compounds W8, W9, and W8a−W8p (Scheme 3). Compound 9 was hydrolyzed to give compound 10 in a DCM solution of trifluoroacetic acid (Scheme 3). Then, AP1 was synthesized using compound 10 as the starting material with the same reaction route as that for compound W8. Tuning the Reactivity of the Reactive Group. Normally, compounds with high reactivity can be easily cleared and can generate nonspecific adducts; therefore, it is extremely important to tune the reactivity of the electrophilic warhead to the intended enzyme target to prevent off-target reactivity.54,55 The intrinsic reactivity of this warhead should be sufficient to covalently modify a specific target when the compound is reversibly bound but not susceptible to chemical stability, metabolic problems, or undifferentiated reactions with other proteins.56 The half-life (t1/2) values of the compounds reacting with buffer, dimethyl sulfoxide (DMSO, a weak nucleophilic reagent57), and glutathione (GSH) provide useful information about their stability, electrophilicity, and likelihood of forming reactive intermediates. In this study, the buffer stability of the compounds was characterized by 1H NMR-based kinetic methods in the absence of GSH. The GSH half-life (GSH t1/2) and DMSO half-life (DMSO t1/2) were determined through the first-order-reaction rate constant (kpseudo‑1st). As illustrated in Figure 2B, W1 showed the appropriate stability (>10 000 min) required for the buffer; however, it was not sufficiently stable (DMSO t1/2 = 52.5 min) in the presence of DMSO, which may be related to the weak nucleophilicity of DMSO. The NMR experiment indicated that the sulfoxide group of DMSO presumably attacks the carbonyl group (blue) of W1, and then undergoes a rearrangement to give the adduct and saccharin (Figure S4). These experimental results showed that W1 was unstable in a highly electrophilic environment, rendering it incompatible for incorporation into druglike systems. Therefore, the systematic study of the reactivity of various electrophiles with DMSO, GSH, and the corresponding structure−reactivity relationship (SRR) is useful for the development of novel covalent warheads; thereby, a series of derivatives of W1 were synthesized to tune its reactivity and stability. 9128 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article Scheme 2. Synthesis of Compounds W4−W7a Reagents and conditions: (a) EtOAc, rt, Et3N, 2 h; (b) THF, 0 °C, 20 min; 50 °C, 6 h; rt, 12 h. a Scheme 3. Synthesis of Compounds W8, W9, and W8a−W8qa Reagents and conditions: (a) triphosgene, DCM, rt, Et3N, 2 h; (b) saccharin, EtOAc, reflux, 1 h. (c) CF3COOH, DCM, 0 °C, 12 h. a negative charge particularly when it exhibits relatively high electronegativity or delocalization with respect to the negative charge. Considering the contributions of carbonyls and sulfones to the electron-withdrawing property of the leaving group (saccharin), we attempted to synthesize compounds W2 and W3 by replacing the carbonyl and sulfone of saccharin with CH2 groups to reduce the rate of step 2 (Figure 2A). Surprisingly, compounds W2 and W3 showed no reactivity toward DMSO (t1/2 > 10 000 min) or GSH (t1/2 > 10 000 According to the proposed mechanism shown in Figure 1C, the overall reaction rate of this type of compound with thiol was controlled by the following two factors: the stability of the carbonyl (step 1) and the effectiveness of the leaving group (step 2). To our knowledge, the less partial positive charge on the carbonyl group is conducive to more stabilization of carbonyl; hence, the electron-donating or -withdrawing capacity of substituents attached to the carbonyl carbon is a primary factor affecting carbonyl stabilization. The effectiveness of a leaving group is related to its ability to stabilize a 9129 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article Figure 2. Tuning the reactivity of the electrophiles. (A) Optimization of strategies for modifying the leaving group ability. (C) Optimization of strategies for modifying the stability of the carbonyl moiety. Evaluation of the half-time (t1/2) of the reaction between electrophiles (B) W1−W3 or (D) W4−W9 and stability in buffer (acetone-d6/D2O = 5:1), DMSO, and GSH. The general formulas of compounds W4−W9 are shown in (C). (E) Crystal structure of W8. CCDC 2105 353 contains the supporting crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/. aReaction conditions: electrophile (20 μmol) in buffer at 37 °C. bReaction conditions: electrophile (20 μmol) in DMSO-d6/D2O (5:1) at 37 °C. cMeasurement could not be obtained due to high reactivity. dReaction conditions: electrophile (20 μmol) and GSH (30 μmol) in DMSO-d6/D2O (5:1) at 37 °C. acceptable reactivity profiles (Figures 2D and S8). To further investigate the effect of the nitrogen atoms on reactivity, an Xray structure of W8 was resolved. As shown in Figure 2E, an intramolecular hydrogen bond (with a distance of N(2)− H(2)···O(1) = 0.88 Å) between the hydrogen atom of the amino group and the oxygen atom of the carbonyl group was formed, which remarkably enhanced the stabilization and reduced the reactivity of the carbonyl carbon in W8 upon nucleophilic attack. Therefore, N-acylamino saccharin can be regarded as a potential emerging covalent warhead for use in designing novel covalent drugs. Next, we synthesized the analogues of W8 by replacing the phenyl group with different groups (Table 1), such as propyl (W8a), thiophen-2-yl (W8b), cyclohexyl (W8c), naphthalen1-yl (W8d), and 1-phenylethyl (W8e), to explore the effect of phenyl on the reactivity of W8. Compared with W8, most of these derived compounds were more reactive toward GSH, the thiophen-2-yl- and naphthalen-1-yl-containing compounds exhibit the most reactivity (GSH t1/2 = 21.9, 23.7 min). Furthermore, a library of W8 derivatives, including those formed through the substitution of withdrawing groups (i.e., trifluoromethyl and halogen) and donating groups (i.e., methyl and methoxy), was synthesized to evaluate the effects of the substituents of the benzene ring on the reactivity of these compounds. Similarly, the majority of these compounds showed increased activity (GSH t1/2 = 11.2−47.5 min), except W8k (GSH t1/2 = 67.3 min). Notably, the electron-donating substituents on benzene appeared to reduce the reactivity of the corresponding compounds; for instance, the half-lives of compounds W8j (GSH t1/2 = 47.5 min) and W8k were longer than those of W8f−W8i (GSH t1/2 = 11.2−35.9 min). Taken together, these experiments suggested that the N-acylamino saccharin set represents a modular chemotype amenable to reactivity fine-tuning. min) (Figure 2B). We hypothesized that the disappearance of the W2 and W3 reactivity could be attributed to the lack of the core moduleN-acetylated saccharin; therefore, we retained N-acetylated saccharin in further warhead optimization experiments. Subsequently, electron-donating alkyl substituents (W4 and W5, Figure 2C) were attached to the carbonyl carbon to reduce the partial positive charge on the carbonyl carbon of the side chain, making them less reactive to nucleophilic attack than the parent compound W1. It is noteworthy that W4 (DMSO t1/2 = 203.8 min) and W5 (DMSO t1/2 = 346.5 min) showed lower reactivity toward DMSO than W1 (Figure 2D), suggesting that the electrondonating ability of the alkyl substituents was sufficient to reduce the rate of step 1. Therefore, we incorporated the greater electron-donating alkoxy groups at the X-position (W6 and W7). Compound W6 (DMSO t1/2 = 182.4 min) exhibited a half-life in DMSO similar to that of W4, but W7 failed to react with DMSO, suggesting that the introduction of alkoxy groups was effective in reducing the rate of step 1. However, W7 maintained significantly high reactivity (GSH t1/2 < 1 min) toward GSH and was therefore further optimized through the introduction of a greater electron-donating group. The nitrogen atom in amides is a powerful electron-donating group through resonance. The lone pair of electrons on a nitrogen atom of an amide can form π bonds with a carbonyl group, thereby reducing the reactivity of the carbonyl and inhibiting the free rotation of the C−N bond in the amide. Moreover, nitrogen is less electronegative than oxygen, and the stabilization of delocalized positively charged resonance structures is usually better than that of other acid derivatives. Therefore, to further reduce the reactivity of the N-acetylated saccharin warhead, W8 and W9 were synthesized by the introduction of a nitrogen atom at the X-position. Surprisingly, W8 (GSH t1/2 = 61.3 min) and W9 (GSH t1/2 = 33.6 min) exhibited high stability (DMSO t1/2 > 10 000 min) and 9130 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article Table 1. Half-Life (t1/2, min) of Synthesized Compounds, Disulfiram and Afatinib, in the Presence of pH 7.4 Buffer, DMSO, and Glutathione (GSH) Reaction conditions: electrophile (20 μmol) in buffer at 37 °C. bReaction conditions: electrophile (20 μmol) in DMSO-d6/D2O (5:1) at 37 °C. Reaction conditions: electrophile (20 μmol) and GSH (30 μmol) in DMSO-d6/D2O (5:1) at 37 °C. a c 9131 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc compounds W8a and W8 (GSH t1/2 = 61.3, 46.5 min, Table 1) showed similar reactivity; however, FBPase inhibition of W8a (>500 μM) was remarkably different from that of W8 (3.4 μM), indicating that the N-substituted benzyl group was essential for the covalent inhibition of FBPase. Similar to W8, W8e−W8m possessed an N-substituted benzyl group, thereby showing high inhibitory activity against FBPase. In particular, W8k exhibited the most potent inhibition with a half-maximal inhibitory concentration (IC50) of 1.7 μM, which was similar to that of disulfiram (1.5 μM). This evidence indicated that the π−π stacking interaction between FBPase and N-acylamino saccharin inhibitors may be crucial for FBPase inhibition. As illustrated in Figure S7, the NMR experiments revealed that the reaction mechanism of W8 with GSH was similar to that of W1, with the formation of N-benzylthiocarbamate from W8 and GSH accompanied by the departure of saccharin; we thus proposed an SN-based mechanism (Figure 3A) for the modification of cysteines in proteins by W8. The liquid chromatography−mass spectrometry (LC−MS) experiments of representative compound W8 and FBPase were carried out in Figure 3B to confirm this reaction mechanism. W8 was first preincubated with FBPase for 10 min and analyzed by mass spectrometry for identification of potential protein−ligand adducts (MS). As illustrated in Figure 3B, peptides (133.05 Da) of C128 covalently bound with part (N-benzyl carboxamide fragment, W8-2) of W8 can be observed, demonstrating that W8 could covalently modify C128 of FBPase. In addition, only the C128A mutation led to a significant increase (over 147-fold, Table S2) in the IC50 of W8, compared to that of wild-type (WT) FBPase. This finding provides new evidence for the confirmation of the newly discovered reaction mechanism (Figure 1C) of N-acylamino saccharin derivatives with cysteine in enzymes. Site-directed mutation assays and LC−MS data showed that, as in our previous studies,47 C128 is extremely important for the covalent reaction of N-acylamino saccharin derivatives with FBPase. Furthermore, the cocrystal structure of the FBPase bound with W8 (PDB ID: 7WJV, Figures 3C and S10) was resolved to elaborate the reaction mechanism of N-acylamino saccharin derivatives and C128 in FBPase. The electron density map of the allosteric site indicated that the N-benzyl carboxamide fragment (W8-2) was close to the C128 residue, and C−S bond formation was unambiguously confirmed by the electron density in chain b (Figure 3C). These results confirm that, similar to W1, W8 covalently modifies the cysteine in FBPase through the newly discovered reaction mechanism, as shown in Figure 1C. The alignment of FBPase and W8 cocrystal structures revealed that W8-2 occupied the allosteric site of FBPase with a binding mode similar to that of 214b (PDB ID: 6LS5).47 Notably, the benzene ring of W8-2 also formed two weakly π−π stacking interactions with the phenyl of residue Y258 and imidazolyl ring of H253 (with the distances of 4.7 and 4.2 Å, respectively)61 and a cation−π stacking interaction with the guanidine group of residue R254 (with a center distance of 5.5 Å),62 demonstrating the importance of the aromatic ring in the activity of the hit compounds. To identify the central and important role of these three residues in covalent inhibitory regulation when FBPase was covalently modified by W8-2, the IC50 values of W8 against Y258A, H253A, and R254A mutants were determined systematically (Table S3). As listed in Table S3, R254A mutation led to a remarkable increase (21.6-fold) in the IC50 of W8, compared To evaluate the chemical stability of N-acylamino saccharin compared with that of commercially available covalent drugs that are known to be covalently bound to cysteine residues, the half-lives of disulfiram and afatinib toward GSH were determined following the aforementioned protocol. Disulfiram, a well-known antialcoholism58 and a first-generation covalent inhibitor of FBPase with a disulfide bond,59 yielded a GSH t1/2 value of 32.8 min. Afatinib, an irreversible tyrosine kinase inhibitor with acrylamide,60 showed lower GSH reactivity (GSH t1/2 = 88.8 min). Compared with disulfiram, compounds W8b, W8d, W8g−W8i, and W8n−W8p exhibited much faster reaction rates (GSH t1/2 = 11.2−29.5 min), while W8e, W8f, W8l, W8m, and W8q (GSH t1/2 = 35.9−31.4 min) showed similar reactivities. In addition, W8, W8a, W8c, W8j, and W8k showed much slower reaction rates (GSH t1/2 = 46.5−67.3 min) than disulfiram. To gain a deeper mechanistic understanding of the reactivity of N-acylamino saccharin with biological thiols and to explain the differences in reactivity between similarly substituted analogues, we calculated the lowest unoccupied molecular orbital (LUMO) energies by the density functional theory (DFT) method, and the results are listed in Table 1. The increase in reactivity observed when going from W8a, W8d, and W8k to W8h, W8i, and W8n can be explained by differences in the electronic factors. Incorporation of an electron-withdrawing group into the aromatic ring (W8h, W8i, W8n) resulted in the lowering of calculated LUMO energies, thereby exhibiting high reactivity. Notably, the strong correlation (R2 = 0.5936, Figure S5) between GSH t1/2 and LUMO energies suggests that the Nacetylated saccharin-based warheads have tunable reactivity by the introduction of electron-withdrawing groups into the aromatic ring. In brief, W8 analogues with moderate to weak reactivity (GSH t1/2 = 31.4−67.3 min) are suitable for the development of target-specific covalent inhibitors. The tunable GSH reactivity of N-acylamino saccharin as a novel covalent warhead establishes a progressively stronger foundation for its further application. Utilization of N-Acylamino Saccharin Warheads in FBPase Inhibitor Scaffolds. Targeting FBPase is an emerging approach for diabetes therapy, and the C128 site on FBPase has recently been proven to be a highly promising strategy for designing drugs with hypoglycemic effects in vivo and in vitro.47 To explore the application of these N-acylamino saccharin derivatives in a druglike setting, eight compounds (Table 2) with low reactivity toward GSH (t1/2 > 30 min) were chosen to evaluate their FBPase inhibitory ability, binding mode, selectivity, and in vivo effect. As listed in Table 2, Table 2. FBPase Inhibitory Activities (IC50 μM) of W8 Analogues compounds IC50 (μM)a W8 W8a W8c W8e W8f W8j W8k W8m disulfiram 3.4 ± 0.5 >500 16.4 ± 1.0 8.7 ± 1.5 4.2 ± 0.1 3.6 ± 0.5 1.7 ± 0.2 1.9 ± 0.5 1.5 ± 0.3 Article a IC50 values were calculated from the FBPase activity assay. 9132 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article Figure 3. (A) Proposed mechanism for covalent modification of FBPase by W8. (B) Mass spectra of FBPase preincubated with W8-2 (predicted Δm, 133.05 Da; found Δm, 133.05 Da). (C) X-ray cocrystal structure of W8-2 and FBPase (PDB ID: 7WJV) (2Fo−Fc omit map contoured at 0.8σ). listed in Table S5, all six compounds exhibited high selectivity (>25-fold) for ALDOs (ALODA, ALDOB, and ALDOC), and three compounds (W8, W8k, and W8m) displayed high GAPDH selectivity (>25-fold). In comparison, W8f showed moderate GAPDH selectivity (16.3-fold). No remarkable GAPDH selectivity was observed for W8e (1.3-fold) and W8j (3.6-fold), as shown in Table S5. Taken together, compounds W8 and W8k exhibited not only high ALDO selectivity (>40-fold) but also high GAPDH selectivity (>27fold). Furthermore, we used the known multitarget covalent inhibitor disulfiram as the control and determined its inhibitory activity against four enzymes. As shown in Tables 1 and 2, the reactivity and inhibitory activity against FBPase of disulfiram (GSH t1/2 = 32.8 min, IC50 = 1.5 μM) were similar to those of W8f (GSH t1/2 = 35.9 min, IC50 = 4.2 μM). However, inhibitory activities of disulfiram against four enzymes (ALDOA, ALDOB, ALDOC, GAPDH) were almost higher than those of W8f (Table S5), indicating that the Nacylamino saccharin warhead provided in this work has some selectivity over disulfiram. to that of WT FBPase, but Y258A and H253A mutations show similar inhibition to WT FBPase. These results indicated that the π−π stacking interactions between W8 and Y258/H253 were weakly possible due to the longer distances between them. In comparison, the cation−π stacking interaction between W8 and H253 was nontrivial for the covalent binding of W8 against FBPase. In conclusion, these findings not only confirmed the covalent binding of W8 to C128 but also more clearly showed the mechanism of the reaction between Nacylamino saccharin derivatives and cysteine residues. Target selectivity is a key but challenging issue in the design of safe and effective covalent ligands due to concerns about the formation of nonspecific or untargeted adducts that lead to potential toxicity. To evaluate the selectivity of the Nacylamino saccharin warhead on the targets in the glucose metabolic pathway, the inhibitory activities of hit compounds (i.e., W8, W8e, W8f, W8j, W8k, and W8m) against several essential targets in the glucose metabolic pathway (i.e., ALDOA, ALDOB, ALDOC, and glyceraldehyde 3-phosphate dehydrogenase (GAPDH)) were evaluated (Table S5). As 9133 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article Figure 4. Effect of a single administration of compound W8 on blood glucose in 7−9 week-old ICR mice. (A) Blood glucose changes and (B) AUC of blood glucose between 0 and 6 h in 12 h fasted ICR mice after oral administration of compound W8 (n = 5−6 in each group). (C) Blood glucose responses to an oral glucose challenge (2 g/kg) and (D) AUC of blood glucose between 0 and 2 h in 12 h fasted ICR mice. Compound W8 was intraperitoneally administered 1 h prior to the oral glucose challenge. (E) Effects of W8 on glucose output in primary rat hepatocytes after overnight serum starvation treatment (LG-DMEM) at concentrations of 50−300 μM. (F) Relative protein expression of primary rat hepatocytes in physical conditions after treatment by compounds W8 at concentrations above 50−300 μM. Data are presented as the mean ± standard error of the mean (SEM) (*P < 0.05, **P < 0.01, ***P < 0.001 vs vehicle; Student’s t-test). Metformin, Met. to 130 kDa. In particular, 2 μM AP1 could predominantly label a 35 kDa protein of the hepatic LO2 cell, as illustrated in Figure S11C, indicating the better selectivity of AP1 in covalent binding with proteins of the LO2 cell. These experimental results further suggested that the N-acylamino saccharin warhead exhibited good selectivity in the LO2 cell. Glucose Reduction in ICR Mice. Our aforementioned experiments demonstrated that N-acylamino saccharin derivatives covalently bound to the C128 site of FBPase to inhibit FBPase enzymatic activity, which encouraged us to further evaluate their glucose-lowering effects in vivo and to investigate the applicability of these covalent warheads in promoting hypoglycemia. Considering their superior FBPase inhibition, compounds W8 and W8k were chosen to evaluate their potency in blood glucose management and the inhibition of glucose output in mouse primary hepatocytes. First, ICR mice were intraperitoneally administered 30 mg/kg saccharin, W8, or W8k; and then, blood glucose was measured during the following 0−6 h period. As shown in Figure S12A,B, 30 mg/kg W8 exhibited an apparent glucose-lowering effect at 2 and 6 h. In comparison, saccharin administered at a concentration of 30 mg/kg, the dose equivalent to that of W8, upon covalently reacting with FBPase, reduced the glucose level, showing no significantly glucose-lowering effect during the 0−6 h period. These results preliminarily indicated that W8 but not saccharin exerted a glucose-lowering effect on the mice model because W8 could covalently bind to FBPase. Notably, 30 mg/kg W8k failed to display a hypoglycemic effect, which may be due to its poor water solubility, which makes it difficult to effectively target hepatic FBPase in vivo. In addition, the inhibitory activities of compounds W8 and W8k against three typical kinases (Bruton’s tyrosine kinase (BTK), epidermal growth factor receptor (EGFR), and Janus kinase 3 (JAK3)), which are common drug targets for cysteine covalent inhibition, were analyzed to assess the selectivity of this type of novel covalent warhead. As shown in Table S6, compounds W8 and W8k exhibited very low inhibition activities in BTK/EGFR/JAK3 (300 μM inhibition rate is less than 50%), indicating that they existed more than 100-fold selectivity between FBPase and BTK/EGFR/JAK3. Therefore, these two N-acylamino saccharin compounds showed a high probability of usefulness in the design of covalent inhibitors against FBPase. Activity-based protein profiling (ABPP) is one of the chemical proteomic approaches that use small-molecule chemical probes to understand the interaction mechanisms between compounds and targets, which can be used to identify the protein targets of small molecules and even the active sites of target proteins. Thus, to assess the proteomic reactivity of compound W8, we designed and synthesized the ABPP probe AP1 based on the structure of W8. Iodoacetamide alkyne (IAA), a well-known nonselective covalent probe, was used as the control. ABPP was performed in hepatic LO2 cells or purified FBPase in terms of the workflow shown in Figure S11A. As illustrated in Figure S11B, AP1 could label the purified FBPase of about 35 kDa in a dose-dependent manner. It should be noticed from Figure S11C that 2−8 μM AP1 could selectively label 35 kDa (the molecular weight of FBPase) and 40 kDa proteins of the hepatic LO2 cell; in comparison, 2 μM IAA could label almost all proteins from 15 9134 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article warheads in drug molecules. The weak reactivity and better target selectivity of N-acylamino saccharin warheads were successfully exploited for the development of covalent inhibitors, and the utility of these novel electrophilic warheads as chemical biological probes was demonstrated through the covalent modification of FBPase, a promising target associated with cancer and type 2 diabetes. The cocrystal structure of compound W8 reacting with FBPase unexpectedly revealed that the N-acylamino saccharin moiety was an electrophile warhead that covalently modified the noncatalytic C128 site of FBPase, and the released saccharin was observed to be nearby the C128 site, suggesting a previously unrecognized covalent reaction mechanism of saccharin derivatives. Moreover, a cation−π stacking interaction between the benzene ring of W8 and the guanidine group of surrounding residue R254 was nontrivial for the covalently binding of W8 against FBPase. Notably, treatment with compounds W8 reduced blood glucose in an ICR mice model, as well as led to dosedependent (50−300 μM) inhibition of glucose production under physiological conditions (P < 0.001 vs vehicle) with low toxicity. Overall, this type of novel warhead with desirable reactivity and stability profiles is easily synthesized and, therefore, is an appropriate supplement to the current repertoire of cysteine covalent modifiers that lack some of the limitations generally associated with established moieties. Notably, a product of this novel covalent reaction is saccharin or a saccharin derivative, which has been proven to be a carbonic anhydrase (CA) inhibitor, hence alternatively suggesting a new strategy for leveraging the protective effect and targeted release of saccharin drugs in the future. Second, ICR mice were also intraperitoneally administered 30 and 10 mg/kg W8 for a detailed evaluation of its hypoglycemic effect. As illustrated in Figure 4A,B, 10 and 30 mg/kg W8 showed significant glucose-lowering effects at 2 h, and the area under the curve (AUC)0−6h analysis revealed that 10 and 30 mg/kg W8 resulted in 15.3% and 42.4% reduction in blood glucose, respectively. Third, an oral glucose tolerance test (OGTT) was performed to explore the glucose tolerance capacity of the lead compound. Figure 4C,D shows that the blood glucose in the control group dramatically increased to 13.34 mM 0.25 h after oral 2 g/kg glucose intake, while the 10 and 30 mg/kg treated groups exhibited a significant suppressive effect on blood glucose. The AUC0−2h analysis also suggested that the 10 and 30 mg/kg treated groups showed reductions in glucose of 20.2% and 17.7%, respectively. In summary, these in vivo experiments clearly demonstrated that W8, an N-acylamino saccharin derivative, exerted a glucose-lowering effect on ICR mice. Inhibition of Glucose Output in Mouse Primary Hepatocytes. FBPase is a gatekeeper enzyme in the hepatic gluconeogenesis pathway (glucose is produced from alanine, glycerol, and lactic acid), which mainly accounts for endogenous glucose production. To explore whether the in vivo glucose-lowering effect of W8 resulted from the inhibition of FBPase, the effects of W8 on gluconeogenesis glucose output by mouse primary hepatocytes were determined. As illustrated in Figure 4E, W8 exhibited remarkable dosedependent glucose-lowering effect on mouse primary hepatocytes at concentrations from 50 to 300 μM, and a cellular viability assay showed that W8 induced no significant toxicity in hepatocytes (Figure 4F), suggesting that the glucoselowering effect of W8 in hepatocytes was reliable. Furthermore, we also investigated the glucose-lowering effect of saccharin in hepatocytes. Saccharin administered at 50 and 100 μM did not lead to an apparent glucose-lowering effect, while the same dose of W8 showed an effect that was consistent with our in vivo results (Figure S12C). These results indicated that the glucose-lowering effect of W8 resulted from the inhibition of the hepatic gluconeogenesis pathway. Pharmacokinetic Properties of Compound W8 and Saccharin. The concentrations of plasma W8 and its leaving group saccharin at different times (0−24 h) were determined (Tables S9 and S10), and the pharmacokinetic properties of saccharin were evaluated in vivo. It could be seen in Table S9 that a little of W8 was detected within 6 h, but a large amount of its leaving group saccharin was detected in plasma at 0.25 h (Table S10), indicating that W8 could react rapidly to generate saccharin in vivo (Figure S15). Saccharin is further metabolized in vivo, and its elimination half-live (T1/2), time to reach maximum plasma concentration (Tmax), maximum plasma concentration (Cmax), and area under the curve (AUC(0−t)) were 1.11 h, 0.25 h, 29 922.33 ng/mL, and 14 034.38 h·ng/mL, respectively (Table S11). ■ EXPERIMENTAL SECTION Chemistry. Common reagents and solvents were purchased from commercial suppliers and used without further purification unless otherwise stated. Reaction progress was monitored using analytical thin-layer chromatography (TLC) on precoated silica gel GF254 plates (QingdaoChem), and spots were detected under UV light (254 and 365 nm). Compounds were purified with flash column chromatography with silica gel and particle size of 48−74 μM (Macklin) as the stationary phase and petroleum ether/ethyl acetate mixture as the eluent system. 1 H and 13C spectra were obtained on a Bruker AV-600 NMR instrument (Bruker, Karlsruhe, Germany) using deuterated solvents (DMSO-d6, CDCl3, acetone-d6, D2O), and 19F NMR was recorded with a Bruker AMX 400 spectrometer with DMSO-d6 and with tetramethylsilane (TMS) as the internal standard. Chemical shifts are expressed in ppm relative to DMSO-d6, CDCl3, acetone-d6 or D2O (2.50/7.26/2.15/4.81 for 1H; 39.52/77.16/29.92 and 206.68 for 13C) with TMS used as the internal standard. The following abbreviations for multiplicity were used: s = singlet, d = doublet, t = triplet, m = multiplet, dd = double doublet, br = broad. High-resolution mass spectrometry (HRMS) data were obtained by electron ionization (EI) using a Waters GCT Premier instrument. Compound purity was determined by high-performance liquid chromatography (HPLC) chromatograms acquired on a DIONEX UltiMate 3000. Analyses were performed using a Thermo Fisher Scientific 120 C18 column (4.6 mm × 250 mm, 5 μm) and acetonitrile for 10 min. Detection was measured at 254 nm, and the average peak area was used to determine purity. All of the compounds were determined to be >95% pure. General Procedure for Synthesizing Compound 2. Saccharin (0.458 g, 2.5 mmol) was added to a cold solution of lithium aluminum hydride (0.19 g) in THF (30 mL) that had been maintained at 0 °C in an external ice bath. The reaction mixture was allowed to reach ambient temperature and stirred for 1 h. The reaction was quenched with the addition of water and 2.5 M aqueous sulfuric acid. The mixture was filtered through Celite and washed with ■ CONCLUSIONS In this study, the discovery of new saccharin moieties capable of covalent reactions with cysteine was reported. Exploration of reactive saccharin derivatives led to the identification of Nacylamino saccharin moieties capable of electrophilic cysteine capture, as determined by combining 1H NMR, site-directed mutagenesis, and protein LC−MS, which hold potential applications as novel chemical biological probes and as new 9135 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Article 2H), 7.35 (t, J = 7.2 Hz, 1H), 5.49 (s, 2H). 13C NMR (151 MHz, CDCl3): δ 151.13, 142.20, 132.46, 131.55, 130.17, 129.10, 124.04, 123.94, 123.48, 121.50, 120.85, 116.54, 65.05. N-Benzyl-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8). Compound W8 was synthesized via the method described by Carnaroglio et al.63 Compound 8 (0.266 g, 2 mmol) was added to a solution of saccharin (0.366 g, 2 mmol) in ethyl acetate (30 mL) and refluxed for 1 h. Then, it was cooled and concentrated in vacuo, the residue was washed with aqueous acetone solution (50% acetone in water), and the target compound W8 was obtained as a white powder (0.632 g, 1.5 mmol, 75% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.66 (s, 1H), 8.32 (s, 1H), 8.19 (d, J = 7.1 Hz, 1H), 8.12 (d, J = 6.5 Hz, 1H), 8.04 (s, 1H), 7.37 (s, 4H), 7.28 (s, 1H), 4.50 (s, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.28, 148.33, 138.67, 137.56, 137.39, 135.86, 128.83, 127.85, 127.60, 126.33, 124.95, 121.89, 43.56. HRMS (ESI) m/z: calcd for C15H12N2O4S [M + H]+, 317.0591; found, 317.0583. 3-Oxo-N-(1-phenylethyl)benzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W9). Compound W9 was synthesized via the same route as that used for compound W8 as a white powder (0.429 g, 1.30 mmol, 52% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.32 (d, J = 7.2 Hz, 1H), 8.18 (d, J = 7.7 Hz, 2H), 8.13 (d, J = 7.0 Hz, 1H), 8.04 (d, J = 7.2 Hz, 1H), 7.32 (d, J = 6.7 Hz, 2H), 7.28 (s, 2H), 7.23 (s, 1H), 3.53 (d, J = 3.7 Hz, 2H), 2.87 (s, 2H). 13C NMR (151 MHz, DMSOd6): δ 158.83, 147.64, 138.62, 137.03, 136.91, 135.33, 128.58, 128.39, 126.26, 125.83, 124.27, 121.37, 41.06, 34.69. HRMS (ESI) m/z: calcd for C16H14N2O4S [M + H]+, 331.0747; found, 331.0741. 3-Oxo-N-propylbenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8a). Compound W8a was synthesized via the same route as that used for compound W8 as a white powder (0.368 g, 1.38 mmol, 55% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.33 (d, J = 7.7 Hz, 1H), 8.22-8.15 (m, 2H), 8.12 (t, J = 7.5 Hz, 1H), 8.04 (t, J = 7.5 Hz, 1H), 3.25 (dd, J = 12.8, 6.4 Hz, 2H), 1.56 (dd, J = 14.3, 7.1 Hz, 2H), 0.90 (t, J = 7.3 Hz, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.33, 148.08, 137.55, 137.38, 135.85, 126.29, 124.90, 121.88, 41.73, 22.53, 11.54. HRMS (ESI) m/z: calcd for C11H12N2O4S [M + H]+, 269.0591; found, 269.0585. 3-Oxo-N-(thiophen-2-ylmethyl)benzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8b). Compound W8b was synthesized via the same route as that used for compound W8 as a white powder (0.483 g, 1.50 mmol, 60% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.71 (t, J = 5.6 Hz, 1H), 8.33 (d, J = 7.7 Hz, 1H), 8.18 (d, J = 7.6 Hz, 1H), 8.12 (t, J = 7.6 Hz, 1H), 8.03 (t, J = 7.6 Hz, 1H), 7.45 (d, J = 4.6 Hz, 1H), 7.09 (d, J = 2.4 Hz, 1H), 7.00 (dd, J = 4.7, 3.6 Hz, 1H), 4.65 (d, J = 5.9 Hz, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.20, 148.18, 141.08, 137.52, 137.41, 135.87, 127.18, 126.91, 126.35, 126.08, 124.87, 121.90, 38.55. HRMS (ESI) m/z: calcd for C13H10N2O4S2 [M + Na]+, 344.9975; found, 344.9974. N-(Cyclohexylmethyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8c). Compound W8c was synthesized via the same route as that used for compound W8 as a white powder (0.483 g, 1.50 mmol, 60% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.30 (s, 1H), 8.16 (d, J = 5.5 Hz, 1H), 8.11 (s, 2H), 8.02 (s, 1H), 3.12 (s, 2H), 1.68 (s, 4H), 1.61 (s, 1H), 1.52 (s, 1H), 1.24-1.08 (m, 3H), 0.93 (d, J = 10.6 Hz, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.43, 148.09, 137.55, 137.37, 135.83, 126.27, 124.92, 121.86, 45.98, 37.62, 30.49, 26.33, 25.72. HRMS (ESI) m/z: calcd for C15H18N2O4S [M + Na]+, 345.0879; found, 345.0874. N-(Naphthalen-1-ylmethyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8d). Compound W8d was synthesized via the same route as that used for compound W8 as a white powder (0.458 g, 1.25 mmol, 50% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.71 (t, J = 5.5 Hz, 1H), 8.32 (d, J = 7.7 Hz, 1H), 8.20 (d, J = 8.3 Hz, 1H), 8.17 (d, J = 7.6 Hz, 1H), 8.11 (t, J = 7.6 Hz, 1H), 8.03 (t, J = 7.6 Hz, 1H), 7.97 (t, J = 10.2 Hz, 1H), 7.89 (d, J = 8.1 Hz, 1H), 7.59 (ddd, J = 22.2, 15.0, 7.1 Hz, 3H), 7.51 (t, J = 7.6 Hz, 1H), 4.98 (d, J = 5.7 Hz, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.37, 148.28, 137.53, 137.39, 135.86, 133.73, 131.06, 129.04, 128.33, 126.90, 126.35, 125.89, 124.96, 123.72, 121.90, 41.62. HRMS (ESI) m/z: calcd for C19H14N2O4S [M + Na]+, 389.0566; found, 389.0556. ethyl acetate. The organic layer was washed with 1 M aqueous sulfuric acid, dried (with anhydrous magnesium sulfate), filtered, and concentrated to give target compound 2, which was a pale-yellow powder (0.36 g, 2.125 mmol, 85% yield). 1H NMR (600 MHz, DMSO-d6): δ 7.81 (d, J = 6.7 Hz, 2H), 7.69 (t, J = 7.2 Hz, 1H), 7.57 (s, 2H), 4.42 (s, 2H). 13C NMR (151 MHz, DMSO-d6): δ 138.29, 136.53, 133.13, 129.37, 125.80, 120.96, 45.33. General Procedure for Synthesizing (1,1-Dioxidobenzo[d]isothiazol-2(3H)-yl)(phenyl)methanone (W2). A solution of benzoyl chloride (0.281 g, 2 mmol) in DCM was added dropwise to a stirred solution of compound 2 (0.338 g, 2 mmol) in DCM with Et3N (0.202 g, 2 mmol), and the solution was stirred at room temperature for 1.5 h. The reaction solution was filtered and concentrated in vacuo, and the residue was recrystallized with dichloromethane to give target compound W2, which was a white powder (0.410 g, 1.50 mmol, 30% yield). 1H NMR (600 MHz, CDCl3): δ 7.87 (d, J = 7.7 Hz, 2H), 7.76 (d, J = 7.8 Hz, 1H), 7.70 (t, J = 7.6 Hz, 1H), 7.62−7.55 (m, 2H), 7.51 (t, J = 7.3 Hz, 3H), 5.20 (s, 2H). 13C NMR (151 MHz, CDCl3): δ 168.72, 134.47, 134.05, 133.94, 132.40, 131.12, 129.62, 128.34, 128.30, 124.83, 121.79, 47.80. HRMS (ESI) m/z: calcd for C14H11NO3S [M + H]+, 274.0532; found, 274.0529. 2-Benzoylisoindolin-1-one (W3). Compound W3 was synthesized via the same route as that used for compound W2 as a faint white powder (0.245 g, 0.9 mmol, 45% yield). 1H NMR (600 MHz, CDCl3): δ 7.87 (d, J = 7.4 Hz, 1H), 7.70 (d, J = 6.8 Hz, 3H), 7.56 (t, J = 8.5 Hz, 2H), 7.52 (t, J = 7.2 Hz, 1H), 7.46 (t, J = 7.2 Hz, 2H), 5.06 (s, 2H). 13C NMR (151 MHz, CDCl3): δ 170.43, 166.87, 141.35, 134.39, 134.17, 131.85, 131.03, 128.75, 128.70, 127.78, 125.34, 123.43, 48.82. 2-(2-Phenylacetyl)benzo[d]isothiazol-3(2H)-one 1,1-Dioxide (W4). A solution of benzoyl chloride (0.281 g, 2 mmol) in ethyl acetate was added dropwise to a stirred solution of saccharin (0.366 g, 2 mmol) in ethyl acetate (20 mL) with Et3N (0.202 g, 2 mmol), and the solution was stirred at room temperature for 2 h. The reaction solution was filtered and concentrated in vacuo, and the residue was recrystallized from dichloromethane to give target compound W4, which was a white powder (0.499 g, 1.66 mmol, 83% yield). 1H NMR (600 MHz, CDCl3): δ 8.14 (d, J = 7.7 Hz, 1H), 7.96 (q, J = 7.4 Hz, 2H), 7.90 (t, J = 7.1 Hz, 1H), 7.34 (d, J = 4.3 Hz, 4H), 7.29 (dd, J = 8.7, 4.3 Hz, 1H), 4.38 (s, 2H). 13C NMR (151 MHz, CDCl3): δ 164.26, 152.63, 133.43, 131.73, 130.13, 126.76, 124.90, 123.92, 122.83, 121.57, 120.15, 116.44, 39.32. HRMS (ESI) m/z: calcd for C15H11NO4S [M + Na]+, 324.0301; found, 324.0295. 2-(3-Phenylpropanoyl)benzo[d]isothiazol-3(2H)-one 1,1-Dioxide (W5). Compound W5 was synthesized via the same route as that used for compound W4 as a white powder (0.511 g, 1.62 mmol, 81% yield). 1H NMR (600 MHz, CDCl3): δ 8.13 (d, J = 7.6 Hz, 1H), 8.02−7.93 (m, 2H), 7.93−7.87 (m, 1H), 7.35−7.24 (m, 4H), 7.22 (t, J = 7.0 Hz, 1H), 3.36 (t, J = 7.6 Hz, 2H), 3.08 (t, J = 7.6 Hz, 2H). 13C NMR (151 MHz, CDCl3): δ 165.78, 152.82, 134.98, 133.36, 131.76, 130.19, 123.84, 123.81, 121.70, 121.52, 120.19, 116.47, 35.13, 24.73. 3-Oxobenzo[d]isothiazole-2(3H)-carboxylate 1,1-Dioxide (W6). Compound W6 was synthesized via the same route as that used for compound W4 as a white powder (0.533 g, 1.76 mmol, 88% yield). 1 H NMR (600 MHz, CDCl3): δ 8.22 (d, J = 7.5 Hz, 1H), 8.02 (s, 2H), 7.95 (d, J = 5.0 Hz, 1H), 7.45 (d, J = 7.4 Hz, 2H), 7.40−7.30 (m, 3H). 13C NMR (151 MHz, CDCl3): δ 155.92, 149.48, 145.35, 137.24, 136.47, 135.04, 129.66, 126.99, 126.40, 125.57, 121.44, 121.16. HRMS (ESI) m/z: calcd for C14H9NO5S [M + H]+, 304.0274; found, 304.0273. Benzyl 3-Oxobenzo[d]isothiazole-2(3H)-carboxylate 1,1-Dioxide (W7). Benzyloxycarbonyl chloride (0.853 g, 5 mmol) was added dropwise to a suspension of saccharin sodium salt (1.025 g, 5 mmol) in ice-cold tetrahydrofuran (50 mL) with stirring. The mixture was maintained at 50 °C for 6 h and at room temperature for an additional 12 h. The insoluble material was removed by filteration, and the filtrate was evaporated. The residue was crystallized from THF to give W7 as a white solid (1.443 g, 4.55 mmol, 91% yield). 1H NMR (600 MHz, CDCl3): δ 8.15 (d, J = 7.7 Hz, 1H), 7.95 (s, 2H), 7.89 (ddd, J = 8.1, 5.4, 3.0 Hz, 1H), 7.50 (d, J = 7.3 Hz, 2H), 7.39 (t, J = 7.3 Hz, 9136 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc 3-Oxo-N-(1-phenylethyl)benzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8e). Compound W8e was synthesized via the same route as that used for compound W8 as a white powder (0.511 g, 1.55 mmol, 62% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.42 (d, J = 6.8 Hz, 1H), 8.32 (d, J = 7.5 Hz, 1H), 8.19 (d, J = 7.3 Hz, 1H), 8.12 (t, J = 7.2 Hz, 1H), 8.05 (d, J = 7.4 Hz, 1H), 7.44 (d, J = 7.0 Hz, 2H), 7.38 (s, 2H), 7.30 (d, J = 6.4 Hz, 1H), 5.08−4.94 (m, 1H), 1.53 (d, J = 6.5 Hz, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.53, 147.27, 143.35, 137.49, 137.44, 135.93, 128.98, 127.73, 126.48, 126.32, 125.00, 121.94, 50.20, 22.67. HRMS (ESI) m/z: calcd for C16H14N2O4S [M + Na]+, 353.0566; found, 353.0565. N-(2-Fluorobenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8f). Compound W8f was synthesized via the same route as that used for compound W8 as a white powder (0.400 g, 1.20 mmol, 48% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.66 (s, 1H), 8.34 (d, J = 7.5 Hz, 1H), 8.21 (d, J = 7.5 Hz, 1H), 8.14 (t, J = 7.2 Hz, 1H), 8.06 (d, J = 7.5 Hz, 1H), 7.48 (d, J = 7.1 Hz, 1H), 7.37 (s, 1H), 7.24 (d, J = 13.7 Hz, 2H), 4.57 (d, J = 5.1 Hz, 2H). 13C NMR (151 MHz, DMSO-d6): δ 161.27, 159.49 (d, J = 48.9 Hz), 148.35, 137.53, 137.42, 135.88, 129.94, 129.78, 126.34, 125.37 (d, J = 13.9 Hz), 124.87, 121.91, 115.69, 115.55, 37.66. HRMS (ESI) m/z: calcd for C15H11N2O4FS [M + H]+, 335.0496; found, 335.0496. N-(2-Chlorobenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8g). Compound W8g was synthesized via the same route as that used for compound W8 as a white powder (0.420 g, 1.20 mmol, 48% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.68 (d, J = 5.4 Hz, 1H), 8.31 (d, J = 7.7 Hz, 1H), 8.18 (d, J = 7.6 Hz, 1H), 8.10 (t, J = 7.6 Hz, 1H), 8.02 (t, J = 7.6 Hz, 1H), 7.50−7.41 (m, 2H), 7.39−7.27 (m, 2H), 4.55 (d, J = 5.9 Hz, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.38, 148.42, 137.53, 137.44, 135.91, 135.65, 132.43, 129.67, 129.48, 129.35, 127.76, 126.37, 124.98, 121.94, 41.59. HRMS (ESI) m/z: calcd for C15H11ClN2O4S [M + H]+, 351.0201; found, 351.0200. N-(2-Bromobenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8h). Compound W8h was synthesized via the same route as that used for compound W8 as a white powder (0.583 g, 1.40 mmol, 56% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.70 (s, 1H), 8.31 (s, 1H), 8.18 (s, 1H), 8.10 (s, 1H), 8.02 (s, 1H), 7.63 (s, 1H), 7.44 (s, 1H), 7.38 (s, 1H), 7.23 (s, 1H), 4.52 (s, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.38, 148.42, 137.51, 137.42, 137.16, 135.89, 132.89, 129.73, 129.34, 128.29, 126.35, 124.96, 122.68, 121.92, 43.99. HRMS (ESI) m/z: calcd for C15H11BrN2O4S [M + Na]+, 416.9515; found, 416.9517. 3-Oxo-N-(2-(trifluoromethyl)benzyl)benzo[d]isothiazole-2(3H)carboxamide 1,1-Dioxide (W8i). Compound W8i was synthesized via the same route as that used for compound W8 as a white powder (0.424 g, 1.13 mmol, 45% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.73 (d, J = 5.3 Hz, 1H), 8.31 (d, J = 7.7 Hz, 1H), 8.18 (d, J = 7.6 Hz, 1H), 8.10 (t, J = 7.6 Hz, 1H), 8.02 (t, J = 7.5 Hz, 1H), 7.73 (d, J = 7.7 Hz, 1H), 7.67 (t, J = 7.5 Hz, 1H), 7.62 (d, J = 7.6 Hz, 1H), 7.49 (t, J = 7.4 Hz, 1H), 4.66 (d, J = 5.5 Hz, 2H). 13C NMR (151 MHz, DMSOd6): δ 159.35, 148.50, 137.51, 137.46, 136.80, 135.92, 133.27, 129.01, 128.17, 126.61, 126.37, 125.75, 124.95, 123.93, 121.95, 120.92. 19F NMR (376 MHz, DMSO-d6): δ −58.98 (s). HRMS (ESI) m/z: calcd for C16H14N2O4F3S [M + Na]+, 407.0284; found, 407.0279. N-(2-Methylbenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8j). Compound W8j was synthesized via the same route as that used for compound W8 as a white powder (0.454 g, 1.28 mmol, 55% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.54 (s, 1H), 8.33 (d, J = 7.7 Hz, 1H), 8.19 (d, J = 7.6 Hz, 1H), 8.12 (t, J = 7.6 Hz, 1H), 8.04 (t, J = 7.6 Hz, 1H), 7.33 (d, J = 3.8 Hz, 1H), 7.20 (s, 3H), 4.49 (d, J = 5.7 Hz, 2H), 2.34 (s, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.37, 148.16, 137.54, 137.40, 136.24, 136.04, 135.89, 135.09, 130.50, 127.92, 127.70, 126.33, 125.01, 121.92, 41.70, 19.10. HRMS (ESI) m/z: calcd for C16H14N2O4S [M + Na]+, 353.0566; found, 353.0559. N-(2-Methoxybenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8k). Compound W8k was synthesized via the same route as that used for compound W8 as a white powder (0.346 g, 1.00 mmol, 40% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.52 (s, Article 1H), 8.29 (d, J = 7.6 Hz, 1H), 8.16 (d, J = 7.6 Hz, 1H), 8.09 (t, J = 7.5 Hz, 1H), 8.00 (t, J = 7.5 Hz, 1H), 7.27 (t, J = 8.3 Hz, 2H), 7.01 (d, J = 7.9 Hz, 1H), 6.91 (t, J = 7.2 Hz, 1H), 4.43 (d, J = 5.5 Hz, 2H), 3.83 (s, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.58, 157.31, 148.22, 137.51, 137.39, 135.86, 129.26, 128.68, 126.36, 125.72, 124.97, 121.89, 120.71, 111.14, 55.88, 39.46. HRMS (ESI) m/z: calcd for C16H14N2O5S [M + Na]+, 369.0516; found, 369.0518. N-(3-Bromobenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8l). Compound W8l was synthesized via the same route as that used for compound W8 as a white powder (0.581 g, 1.48 mmol, 59% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.71 (s, 1H), 8.29 (d, J = 6.2 Hz, 1H), 8.16 (d, J = 6.0 Hz, 1H), 8.08 (s, 1H), 8.00 (s, 1H), 7.56 (s, 1H), 7.44 (s, 1H), 7.34 (s, 1H), 7.29 (s, 1H), 4.45 (s, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.21, 148.44, 141.67, 137.54, 137.37, 135.87, 131.00, 130.58, 130.43, 126.99, 126.34, 125.02, 122.09, 121.90, 43.01. HRMS (ESI) m/z: calcd for C15H11BrN2O4S [M + H]+, 394.9696; found, 394.9693. N-(3-Methoxybenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8m). Compound W8m was synthesized via the same route as that used for compound W8 as a white powder (0.652 g, 1.75 mmol, 70% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.66 (s, 1H), 8.33 (d, J = 7.7 Hz, 1H), 8.19 (d, J = 7.5 Hz, 1H), 8.12 (t, J = 7.4 Hz, 1H), 8.04 (t, J = 7.5 Hz, 1H), 7.28 (t, J = 7.7 Hz, 1H), 6.95 (d, J = 7.9 Hz, 2H), 6.85 (d, J = 7.8 Hz, 1H), 4.47 (d, J = 5.5 Hz, 2H), 3.75 (s, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.75, 159.25, 148.28, 140.20, 137.55, 137.37, 135.84, 129.91, 126.31, 124.96, 121.88, 119.91, 113.58, 112.82, 55.41, 43.49. HRMS (ESI) m/z: calcd for C16H14N2O5S [M + Na]+, 396.0516; found, 396.0510. 3-Oxo-N-(4-(trifluoromethyl)benzyl)benzo[d]isothiazole-2(3H)carboxamide 1,1-Dioxide (W8n). Compound W8n was synthesized via the same route as that used for compound W8 as a white powder (0.610 g, 1.58 mmol, 63% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.81 (s, 1H), 8.33 (d, J = 7.3 Hz, 1H), 8.21 (d, J = 7.3 Hz, 1H), 8.13 (t, J = 7.3 Hz, 1H), 8.05 (t, J = 6.8 Hz, 1H), 7.73 (d, J = 7.4 Hz, 2H), 7.60 (d, J = 7.0 Hz, 2H), 4.58 (d, J = 5.4 Hz, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.21, 148.50, 143.71, 137.55, 137.41, 135.88, 128.48, 128.07 (t, J = 114 Hz), 127.95, 126.33, 125.68, 124.93, 121.91, 43.23. 19F NMR (376 MHz, DMSO-d6): δ −60.83 (s). HRMS (ESI) m/z: calcd for C16H11F3N2O4S [M + H]+, 385.0464; found, 385.0467. N-(4-Bromobenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8o). Compound W8o was synthesized via the same route as that used for compound W8 as a white powder (0.604 g, 1.53 mmol, 61% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.69 (s, 1H), 8.28 (s, 1H), 8.15 (s, 1H), 8.09 (s, 1H), 8.00 (s, 1H), 7.52 (s, 2H), 7.30 (s, 2H), 4.42 (s, 2H). 13C NMR (151 MHz, DMSO-d6): δ 159.22, 148.40, 138.28, 137.54, 137.41, 135.88, 131.68, 130.14, 126.34, 124.93, 121.91, 120.65, 43.01. HRMS (ESI) m/z: calcd for C15H11BrN2O4S [M + H]+, 394.9695; found, 394.9696. N-(4-Methoxybenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8p). Compound W8p was synthesized via the same route as that used for compound W8 as a white powder (0.520 g, 1.5 mmol, 60% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.56 (s, 1H), 8.29 (d, J = 7.6 Hz, 1H), 8.15 (d, J = 7.6 Hz, 1H), 8.09 (t, J = 7.5 Hz, 1H), 8.00 (t, J = 7.5 Hz, 1H), 7.28 (d, J = 8.2 Hz, 2H), 6.89 (d, J = 8.3 Hz, 2H), 4.38 (d, J = 5.5 Hz, 2H), 3.71 (s, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.29, 158.92, 148.20, 137.55, 137.39, 135.86, 130.57, 129.43, 126.32, 124.93, 121.89, 114.22, 55.50, 43.07. HRMS (ESI) m/z: calcd for C16H14N2O5S [M + Na]+, 369.0510; found, 369.0516. N-(4-Methylbenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (W8q). Compound W8q was synthesized via the same route as that used for compound W8 as a white powder (0.538 g, 1.63 mmol, 65% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.62 (s, 1H), 8.32 (d, J = 7.2 Hz, 1H), 8.18 (d, J = 7.1 Hz, 1H), 8.11 (d, J = 6.9 Hz, 1H), 8.04 (d, J = 7.0 Hz, 1H), 7.27 (d, J = 6.7 Hz, 2H), 7.17 (d, J = 6.7 Hz, 2H), 4.44 (d, J = 4.3 Hz, 2H), 2.29 (s, 3H). 13C NMR (151 MHz, DMSO-d6): δ 159.29, 148.25, 137.55, 137.40, 136.75, 135.87, 135.61, 129.38, 127.90, 126.33, 124.94, 121.90, 43.34, 21.14. 9137 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc HRMS (ESI) m/z: calcd for C16H14N2O4S [M + H]+, 331.0747; found, 331.0743. N-(4-Ethynylbenzyl)-3-oxobenzo[d]isothiazole-2(3H)-carboxamide 1,1-Dioxide (AP1). tert-Butyl (4-ethynylbenzyl)carbamate (9, 0.462 g, 2 mmol) was resolved in DCM (20 mL), then CF3COOH (2 mL) was added, and the solution was stirred at 0 °C for 12 h. Then, it was concentrated in vacuo, the residue was resolved in saturated sodium carbonate aqueous solution and extracted with ethyl acetate (3 × 100 mL). The organic layer was dried with anhydrous magnesium sulfate, filtered, and concentrated to give compound (4ethynylphenyl)methanamine (10). Then, compound AP1 was synthesized via the same route as that used for compound W8 as a white powder (0.306 g, 0.9 mmol, 45% yield). 1H NMR (600 MHz, DMSO-d6): δ 8.73 (s, 1H), 8.33 (d, J = 7.3 Hz, 1H), 8.20 (d, J = 7.0 Hz, 1H), 8.12 (d, J = 7.1 Hz, 1H), 8.05 (d, J = 7.2 Hz, 1H), 7.48 (d, J = 7.3 Hz, 2H), 7.39 (d, J = 7.3 Hz, 2H), 4.51 (d, J = 4.7 Hz, 2H), 4.18 (s, 1H). 13C NMR (151 MHz, DMSO-d6): δ 159.23, 148.41, 139.76, 137.54, 137.39, 135.86, 132.16, 128.06, 126.32, 124.93, 121.89, 120.89, 83.77, 81.17, 43.31. HRMS (ESI) m/z: calcd for C17H12N2O4S [M + Na]+, 363.0415; found, 363.0415. NMR Method to Determine Kinetic Parameters. The analysis method from reported procedures was adopted with modifications.64,65 The reaction temperature was 37 °C in all cases. A 500 μL aliquot of the 40 mM test compound was placed in an NMR tube with 5 mm outside diameter (OD). A freshly prepared solution (100 μL) of nucleophile stock (30 μmol) in buffer (300 mM, pH 7.4, D2O) was transferred to the NMR tube that contained reagents and was quickly inverted several times to aid mixing and dissolution; thus, the reaction was initiated. The thoroughly mixed solution was inserted into the NMR machine cavity, and the acquisition of 1H NMR data was immediately initiated (TMS was used as an internal standard for the NMR analysis as it did not interfere with the analyses). The pseudo-first-order rate constants were determined by plotting the natural log of the electrophile/internal standard ratio as a function of time, as defined by the area of a given resonance OD an electrophile or internal standard vs time. The negative slope of the straight line is the pseudo-first-order rate constant. inhibition assays, mass spectrometry studies, protein crystallization, etc. In Vitro Biological assays. For the human liver FBPase inhibition assay, the enzymatic activities of Hu-FBPase were measured by a colorimetric assay based on the detection of inorganic phosphate hydrolyzed from FBP, as described in our previous work.67 The released phosphate was quantified in a complex with ammonium molybdate and malachite green by spectrophotometry. To calculate the product formation at the micromolar level, calibration curves were generated using a standard KH2PO4 solution in the range of 2−60 μM. The absorbance of the reaction mixture was measured at 620 nm with a spectrophotometer (SpectraMax M5, Molecular Devices). To determine the corresponding inhibitor constants (IC50 values), initial rate data of the saturating substrate, fixed effector, and systematically varied inhibitor concentrations were fit to the Hill equation: V = V0 − (V0 − V∞)/[(IC50/I)n + 1], where V, V0, and V∞ are the velocity, maximum velocity (at I = 0), and limiting velocity (at I saturation), respectively; n is the Hill coefficient associated with the inhibitor; and IC50 is the concentration of inhibitor to reach a 50% inhibition rate. All kinetic data were fit to a growth/sigmoidal model with Origin 7.5 software. Selectivity Evaluation of Some Saccharin Derivatives. The pET28a expression vector containing the human GAPDH gene was cloned. The enzyme was expressed in E. coli BL21 (DE3) and purified to homogeneity as previously described.68 In vitro recombinant human GAPDH activity was measured by spectrophotometry as described by Kornberg with slight modifications.69 Assays were performed with 10 mM sodium pyrophosphate buffer (pH 8.5) in 96well plates. First, 495 μL of recombinant GAPDH (0.1 mg/mL final concentration) was incubated with 5 μL of test compounds for 30 min. Then, the enzymatic activity was measured with a microplate reader spectrophotometer (SpectraMax M5, Molecular Devices), with absorbance at 340 nm, indicating the reduction of NAD+. The assay was performed at 37 °C. An additional 200 μL of reaction mixture containing sodium arsenate, 4 mM NAD+, and 12 mM glyceraldehyde 3-phosphate (G3P) was then rapidly added to each well to start the reaction. The absorbance was measured at 340 nm 4 min after reaction initiation. A nicotinamide adenine dinucleotide (NADH)-linked enzymatic assay was performed to measure the inhibitory activity of compounds against aldolase.70 Commercial preparations of glycerol 3-phosphate dehydrogenase (GPDH) from rabbit muscle and triosephosphate isomerase (TPI) from rabbit muscle, both obtained from Sigma, were used. Recombinant aldolase (ALDOA/ALDOB/ALDOC) was mixed with serial dilutions of each compound, NADH (0.41 mM) and triosephosphate isomerase (TPI) (0.0025 U/μL, Sigma), in 412 μL of assay buffer (0.1 M Tris, pH 7.4 and 0.2 M potassium-acetate) and incubated for 3 min at 37 °C. The reaction was initiated by adding 4.2 μL of FBP (100 mM) and 4.2 μL of GAPDH (0.0625 U/μL, Sigma). The decrease in NADH absorbance at 340 nm was measured every 30 s for 6 min with a spectrophotometer (SpectraMax M5, Molecular Devices). Initial velocities of reactions with the same compounds used in combination with DMSO were calculated and used to generate IC50 curves. The kinase assays were carried out as described previously.71 All of the enzymatic reactions were conducted at 30 °C for 40 min. The 50 μL reaction mixture contains 40 mM Tris, pH 7.4, 10 mM MgCl2, 0.1 mg/mL bovine serum albumin (BSA), 1 mM dithiothreitol (DTT), 10 μM adenosine 5′-triphosphate (ATP), kinase, and the substrate. The compounds were diluted in 10% DMSO, and 5 μL of the dilution was added to a 50 μL reaction so that the final concentration of DMSO is 1% in all of the reactions. The assay was performed using the Kinase-Glo Plus luminescence kinase assay kit. It measures kinase activity by quantitating the amount of ATP remaining in the solution following a kinase reaction. The luminescent signal from the assay is correlated with the amount of ATP present and is inversely correlated with the amount of kinase activity. Preparation of Cell Lysates. Hepatic LO2 cells were maintained in Roswell Park Memorial Institute (RPMI)-1640 medium (Procell) supplemented with 10% fetal bovine serum (FBS) and 1% antibiotics. ln([electrophile]) = − k pseudo 1st × t + ln([electrophile0]) t1/2 = Article 0.693 (60×k pseudo 1st) Protein Expression and Purification. The cDNA of Hu-FBPase (GenBank: D26055.1) was cloned into an EX-C0133-B01 vector. To obtain the purified protein, an eight amino acid sequence (SYYHHHHHH) was added to the N-terminus of Hu-FBPase. The plasmid was then transformed into BL21 (DE3) cells for protein expression and purified with a HisTrap_FF_5 mL [Global] column with the standard Ä KTA pure system, as described in our previous work.66 The process for purifying hGAPDH was the same as that used to purify Hu-FBPase, as described above. The cDNA of ALDOA (GenBank: CR541880.1), ALDOB (GenBank: KR711267.1), and ALDOC (GenBank: CR541881.1) was cloned into an pET28a vector. The transformation and purification processes applied to the aldolases were the same as those used for Hu-FBPase, as described above. Site-directed mutagenesis experiments were performed by introducing specific bases into a double-stranded DNA plasmid. Mutant constructs were generated using the two-step polymerase chain reaction (PCR) method. DNA encoding WT Hu-FBPase was cloned into EX-C0133-B01 and used as the template for mutagenesis. Parental methylated and hemimethylated DNA were digested by the NspV and NotI restriction enzymes. Then, the mutant constructs were ligated into a previous plasmid. The plasmids carrying the recombinant mutant were transformed into DH5α competent cells. All of the mutations were confirmed by DNA sequencing. The verified plasmids with mutations were transformed into the Escherichia coli BL21 (DE3) strain cells. The mutant Hu-FBPase proteins were purified in the same manner as WT Hu-FBPase. The eluted protein samples were stored in a solution (10 mM Tris, pH 7.5) for enzymatic 9138 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc The cell lines were grown at 37 °C in a 5% CO2 atmosphere. LO2 cells were grown in the culture medium until 70−80% confluence, then the medium was removed, and the cells were washed twice with cold PBS (10 mL) and lysed with 200 μL of cell lysis buffer for western blot and IP (containing protease inhibitor, Beyotime) analyses. The lysed cells were centrifuged (15 000g, 10 min) at 4 °C. The supernatant was transferred to a separated microfuge tube and stored at −20 °C. After thawing the supernatant on ice, the protein concentration was determined using the bicinchoninic acid (BCA) protein assay kit (Beyotime) and adjusted to 2 mg/mL by dilution with PBS. In Vitro Labeling of LO2 Cell Lysate Proteomes. The cell lysate (48 μL containing 2 mg/mL proteins) was treated with a series of concentrations of AP1 (1 μL) at 37 °C for 2 h. The solution was subjected to a copper-catalyzed azide−alkyne cycloaddition (CuAAC) reaction with 0.5 mM biotin-PEG3-azide, 0.5 mM sodium ascorbate, 0.5 mM 3,3′,3″-[nitrilotris(methylene-1H-1,2,3-triazole-4,1-diyl)] tri(1-propanol), and 0.5 mM CuSO4. The mixture was incubated at 37 °C for 1 h. After CC, 50 μL of 2× sodium dodecyl sulfate (SDS) loading buffer was added to the mixture to stop the reaction, which was further heated at 95 °C for 10 min. The solution was used for subsequent silver staining and streptavidin blotting. Streptavidin Blot Analysis. Denatured protein samples were resolved by 10% SDS-polyacrylamide gel electrophoresis (PAGE), and then, the separated proteins were transferred to poly(vinylidene difluoride) (PVDF) membranes, which were washed with Trisbuffered saline containing 0.1% Tween 20 (TBST) three times powdered milk for 1.5 h, incubated with streptavidin−horseradish peroxidase (HRP, Beyotime, with 1:5000 dilution in 5% nonfat dried milk TBST solution) overnight at 4 °C for biotin labels on blots. Finally, the membrane was washed with PBST, and signals were detected by ECL western blot detection reagents (Beyotime) using a ChemiDoc XRS+ system (Bio-Rad). Protein Crystallization and Structure Determination. W8 was added to the protein such that the final inhibitor concentration after the subsequent addition of the sample was 10 mM. The sample was concentrated by ultrafiltration to a protein concentration of 8 mg/mL (measured by absorbance at 280 nm). Crystals were obtained using a hanging-drop vapor diffusion method at 18 °C. Crystals grew from a mixture of 1 μL of protein and 1 μL of a well solution containing 1 mM adenosine monophosphate (AMP), 0.1 M Tris (pH = 6.0) and 14% v/v EtOH. Crystals were cryoprotected using a well solution supplemented with 25% glycerol and flash-frozen in liquid nitrogen. X-ray diffraction data were collected at the BL19U1 beamline at the Shanghai Synchrotron Radiation Facility. The statistics related to data are presented in Table S6. The diffraction data were indexed, integrated, and scaled using XDS.72 The structure was resolved by molecular replacement using Phaser,73 and the structure of Hu-FBPase (PDB ID: 5ZWK) was used as the search model.74 There were four molecules in the asymmetric unit. The model was built using Coot70 and refined with PHENIX.75 Atomic restraints were generated for the inhibitor using eLBOW,76 and the model was validated using MolProbity.77 Mass Spectrometry. For precipitation and digestion, proteins were precipitated with precooled acetone. The protein pellet was dried by a SpeedVac for 1−2 min. The pellet was subsequently dissolved in 8 M urea and then diluted ten times with 100 mM Tris− HCl. The protein source was digested overnight with Glu-C at a 1:20 ratio (w/w) (Promega, https://www.promega.com.cn/). The reaction was stopped by adding formic acid, and the peptide solution was desalted with a Monospin C18 column (Shimadzu GL). For the LC-tandem MS (MS/MS) analyses of peptides, the peptide mixture was solubilized in 0.1% formic acid and loaded onto a homemade 30 cm long pulled-tip analytical column (ReproSil-Pur C18 AQ with a 1.9 μm particle size, Dr. Maisch GmbH, 75 μm ID × 360 μm OD) connected to an Easy-nLC1200 UHPLC (Thermo Scientific) for mass spectrometry analysis (Q Exactive Orbitrap mass spectrometer, Thermo Scientific, San Jose, CA). The elution gradients and mobile phases used for peptide separations are constituted as follows: for 0−1 min, 3−6% B; for 1−96 min, 6−30% B; for 96−114 Article min, 30−60% B; for 114−115 min, 60−100% B; for 115−120 min, and for 100−100% B (mobile phase A, 0.1% formic acid in water and mobile phase, phase B, 0.1% formic acid in 80% acetonitrile) at a flow rate of 300 nL/min. Peptides eluted from the LC column were directly electrosprayed into the mass spectrometer with the application of a distal 1.8 kV spray voltage. Survey full-scan MS spectra (from m/z 300 to 1800) were acquired with an Orbitrap analyzer (Q Exactive) at a resolution of r = 70 000 at m/z 400. The top 20 MS/MS events were sequentially generated from the full MS spectrum at a 30% normalized collision energy. The dynamic exclusion time was 10 s. Data Analysis. The acquired MS/MS data were analyzed against a database downloaded from UniProt using PEAKS (version 8.5). To estimate peptide probabilities and false discovery rates accurately, we used a decoy database containing the reversed sequences of all of the proteins and appended it to the target database. Mass tolerances for the precursor ions were set at 20 ppm, and for MS/MS, they were set at 0.02 Da. The specific binding was set as a dynamic modification with a mass shift of 342.00198 at cysteine. Glucose Production Assay. Primary mouse hepatocytes treated overnight with serum starvation were treated with compounds in a DMEM environment for 6 h, and the effect of metformin (1 mM), saccharin (50 or 100 μM), and the test compounds (50−300 μM) on the level of gluconeogenesis in primary hepatocytes was measured by assaying the level of glucose in their culture medium. Animals. Male ICR mice (20−24 g) were obtained from Shanghai JSJ Lab Animal, Ltd. The mice were cultured under a specific pathogen-free environment with a 12 h light−dark cycle, relative humidity of 55−60%, a temperature of 22−24 °C, and free access to water and food. The animal studies were approved by the Animal Care and Use Committee of Central China Normal University (CCNU-IACUC-2021-006). Glucose Reduction in ICR Mice. ICR mice (n = 4 in each group) fasted for 12 h were intraperitoneally administered vehicle (10% ricinus oil in water), metformin (250 mg/kg), or a test compound (30 mg/kg). Food was withheld throughout the study. Blood samples were collected from the tail vein 0, 1, 2, 4, and 6 h after treatment and analyzed by a glucometer (Sanicare). OGTT Was Performed with ICR Mice. ICR mice that had been fasted for 12 h (n = 6 in each group) were intraperitoneally administered compound W8 (10 and 30 mg/kg), metformin (250 mg/kg), or vehicle (10% ricinus oil in water) 1 h before the glucose challenge. Glucose (2 g/kg) was orally administered at 0 h, and blood samples were drawn from the tail vein 0, 0.25, 0.5, 1, 1.5, and 2 h after glucose administration. Plasma glucose was measured using a glucometer (Sanicare). Food was withheld throughout the study. Pharmacokinetic Studies. The pharmacokinetic studies of W8 and its leaving group saccharin were performed by Shanghai Medicilon Inc. ICR mice were fasted for 12 h before intraperitoneal administration of 30 mg/kg W8. Water and food were allowed free access after 4 h of intraperitoneal administration. The blood of ICR mice was collected at different time points (0.25, 0.5, 1, 2, 4, 6, 8, and 24 h), 0.20 mL/time point. Tubes with K2EDTA were used to collect blood samples. All samples were centrifuged at 6800g at 4 °C for 5 min within 0.5 h and then were stored at −70 °C until LC−MS/MS analysis. The data on the plasma concentration were analyzed, and the key pharmacokinetic parameters were obtained. ■ ASSOCIATED CONTENT * Supporting Information sı The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jmedchem.2c00336. 1 H NMR spectra of the products from the reaction of W1 and 1,4-benzenedithiol; reaction mechanism of W1 with GSH or DMSO-d6 and W8 with GSH; reaction modeling and resulting energies at the theoretical level; correlation between GSH t1/2 and LUMO energies; halflife data for reaction of W1 with GSH or DMSO-d6; 9139 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Lixia Wu − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China Jiaqi Liu − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China Chen Su − National Facility for Protein Science in Shanghai, Zhangjiang Lab, Shanghai 201210, China Chao Peng − National Facility for Protein Science in Shanghai, Zhangjiang Lab, Shanghai 201210, China crystal data and structure refinement for W8; crystallography data collection and refinement statistics of FBPase−W8 (7WJV); IC50 of W8 against mutants of FBPase; inhibition of W8 and W8k against BTK/ EGFR/JAK3; LC−MS data of FBPase and W8; ABPPbased proteomic analysis; X-ray cocrystal structures of leaving groupssaccharin and FBPase; effect of compounds on blood glucose in the ICR mice model and glucose output in primary rat hepatocytes; HPLC analysis data; and 1H NMR, 13C NMR, and 19F NMR spectra (PDF) Crystallographic data of W8 (CIF) Structure of FBPase in a complex with W8 (PDB) Molecular formula strings (CSV) Complete contact information is available at: https://pubs.acs.org/10.1021/acs.jmedchem.2c00336 Author Contributions ∥ W.W., H.C., and Y.X. contributed equally to this work. All authors contributed to the writing of the manuscript and have approved the final version of the manuscript. Accession Codes The atomic coordinates and structure factors have been deposited into the RCSB Protein Data Bank with accession number 7WJV. The authors will release the atomic coordinates and experimental data upon article publication. ■ Article Funding This work was supported by the Natural Science Foundation of China (Nos. 22177036, 21877046, and 21572077), the Program for PCSIRT (No. IRT0953), the Guizi Scholarship of CCNU (No. 31101222098), the self-determined research funds of CCNU from the colleges’ basic research and operation of MOE (Nos. CCNU19TS011 and CCNU16A02041), and the support from the Program of Introducing Talents of Discipline to Universities of China (111 Program, B17019) is also appreciated. AUTHOR INFORMATION Corresponding Authors Yanliang Ren − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China; orcid.org/0000-0001-8565-2152; Phone: +86-27-67862022; Email: renyl@ccnu.edu.cn Yunyuan Huang − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China; Shanghai Key Laboratory of New Drug Design, School of Pharmacy, East China University of Science and Technology, Shanghai 200237, China; Phone: +86-2767862022; Email: huangyy@ecust.edu.cn Jian Wan − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China; orcid.org/0000-0003-4172-1392; Phone: +86-2767862022; Email: jianwan@ccnu.edu.cn Notes The authors declare no competing financial interest. ■ ACKNOWLEDGMENTS The authors thank the staff members of the Mass Spectrometry System at the National Facility for Protein Science in Shanghai (NFPS), Zhangjiang Lab, China for providing technical support and assistance in data collection and analysis. The authors thank the staff of the BL18U1 and BL19U1 beamline of the NCPSS at the Shanghai Synchrotron Radiation Facility for assistance in data collection. ■ ABBREVIATIONS USED ABPP, activity-based protein profiling; ADMET, absorption, distribution, metabolism, excretion, and toxicity; AMP, adenosine monophosphate; AUC, area under the curve; BTK, Bruton’s tyrosine kinase; CA, carbonic anhydrase; DCM, dichloromethane; DFT, density functional theory; DMSO, dimethyl sulfoxide; EGFR, epidermal growth factor receptor; EI, electron ionization; Et3N, triethylamine; FDA, Food and Drug Administration; FBP, fructose-1,6-diphosphate; FBPase, fructose-1,6-bisphosphatase; GAPDH, glyceraldehyde 3-phosphate dehydrogenase; G3P, glyceraldehyde 3phosphate; GSH, glutathione; HPLC, high-performance liquid chromatography; HRMS, high-resolution mass spectrometry; IAA, iodoacetamide alkyne; IC50, half-maximal inhibitory concentration; ICR, Institute of Cancer Research; JAK3, Janus kinase 3; LC−MS, liquid chromatography−mass spectrometry; LUMO, lowest unoccupied molecular orbital; Met, metformin; MS, mass spectrometry; NADH, nicotinamide adenine dinucleotide; NAS, noncalorie artificial sweetener; NMR, nuclear magnetic resonance; OD, outside diameter; OGTT, oral glucose tolerance test; QM, quantum mechanical; SN, nucleophilic substitution; SRR, structure− reactivity relationship; TCI, targeted covalent inhibitor; TIM, Authors Wuqiang Wen − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China Hongxuan Cao − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China Yixiang Xu − Shanghai Key Laboratory of New Drug Design, School of Pharmacy, East China University of Science and Technology, Shanghai 200237, China Li Rao − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China; orcid.org/0000-0002-0780-6504 Xubo Shao − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China Han Chen − Key Laboratory of Pesticide & Chemical Biology (CCNU), Ministry of Education, College of Chemistry, Central China Normal University, Wuhan 430079, China 9140 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Noe, M. C.; Obach, R. S.; Philippe, L.; Shanmugasundaram, V.; Shapiro, M. J.; Starr, J.; Stroh, J.; Che, Y. Chemical and computational methods for the characterization of covalent reactive groups for the prospective design of irreversible inhibitors. J. Med. Chem. 2014, 57, 10072−10079. (17) Gersch, M.; Kreuzer, J.; Sieber, S. A. Electrophilic natural products and their biological targets. Nat. Prod. Rep. 2012, 29, 659− 682. (18) Gehringer, M.; Laufer, S. A. Emerging and re-emerging warheads for targeted covalent inhibitors: applications in medicinal chemistry and chemical biology. J. Med. Chem. 2019, 62, 5673−5724. (19) Jackson, P. A.; Widen, J. C.; Harki, D. A.; Brummond, K. M. Covalent modifiers: a chemical perspective on the reactivity of alpha,beta-unsaturated carbonyls with thiols via hetero-michael addition reactions. J. Med. Chem. 2017, 60, 839−885. (20) Shindo, N.; Fuchida, H.; Sato, M.; Watari, K.; Shibata, T.; Kuwata, K.; Miura, C.; Okamoto, K.; Hatsuyama, Y.; Tokunaga, K.; Sakamoto, S.; Morimoto, S.; Abe, Y.; Shiroishi, M.; Caaveiro, J. M. M.; Ueda, T.; Tamura, T.; Matsunaga, N.; Nakao, T.; Koyanagi, S.; Ohdo, S.; Yamaguchi, Y.; Hamachi, I.; Ono, M.; Ojida, A. Selective and reversible modification of kinase cysteines with chlorofluoroacetamides. Nat. Chem. Biol. 2019, 15, 250−258. (21) Brouwer, A. J.; Á lvarez, N. H.; Ciaffoni, A.; van de Langemheen, H.; Liskamp, R. M. Proteasome inhibition by new dual warhead containing peptido vinyl sulfonyl fluorides. Bioorg. Med. Chem. 2016, 24, 3429−3435. (22) Narayanan, A.; Jones, L. H. Sulfonyl fluorides as privileged warheads in chemical biology. Chem. Sci. 2015, 6, 2650−2659. (23) Zambaldo, C.; Vinogradova, E. V.; Qi, X.; Iaconelli, J.; Suciu, R. M.; Koh, M.; Senkane, K.; Chadwick, S. R.; Sanchez, B. B.; Chen, J. S.; Chatterjee, A. K.; Liu, P.; Schultz, P. G.; Cravatt, B. F.; Bollong, M. J. 2-Sulfonylpyridines as tunable, cysteine-reactive electrophiles. J. Am. Chem. Soc. 2020, 142, 8972−8979. (24) Elbrecht, A.; Chen, Y.; Adams, A.; Berger, J.; Griffin, P.; Klatt, T.; Zhang, B.; Menke, J.; Zhou, G.; Smith, R. G.; Moller, D. E. L764406 is a partial agonist of human peroxisome proliferator-activated receptor γ. The role of Cys13 in ligand binding. J. Biol. Chem. 1999, 274, 7913−7922. (25) Shearer, B. G.; Wiethe, R. W.; Ashe, A.; Billin, A. N.; Way, J. M.; Stanley, T. B.; Wagner, C. D.; Xu, R. X.; Leesnitzer, L. M.; Merrihew, R. V.; Shearer, T. W.; Jeune, M. R.; Ulrich, J. C.; Willson, T. M. Identification and characterization of 4-chloro-N-(2-{[5trifluoromethyl)-2-pyridyl]sulfonyl} ethyl)benzamide (GSK3787), a selective and irreversible peroxisome proliferator-activated receptor δ (PPARδ) antagonist. J. Med. Chem. 2010, 53, 1857−1861. (26) Lucas, S. D.; Costa, E.; Guedes, R. C.; Moreira, R. Targeting COPD: advances on low-molecular-weight inhibitors of human neutrophil elastase. Med. Res. Rev. 2013, 33, E73−E101. (27) Hacker, S. M.; Backus, K. M.; Lazear, M. R.; Forli, S.; Correia, B. E.; Cravatt, B. F. Global profiling of lysine reactivity and ligandability in the human proteome. Nat. Chem. 2017, 9, 1181− 1190. (28) Weihrauch, M. R.; Diehl, V. Artificial sweeteners–do they bear a carcinogenic risk? Ann. Oncol. 2004, 15, 1460−1465. (29) Jo, J. H.; Kim, S.; Jeon, T. W.; Jeong, T. C.; Lee, S. Investigation of the regulatory effects of saccharin on cytochrome P450s in male ICR mice. Toxicol. Res. 2017, 33, 25−30. (30) Moeker, J.; Peat, T. S.; Bornaghi, L. F.; Vullo, D.; Supuran, C. T.; Poulsen, S. A. Cyclic secondary sulfonamides: unusually good inhibitors of cancer-related carbonic anhydrase enzymes. J. Med. Chem. 2014, 57, 3522−3531. (31) Morku̅naitė, V.; Baranauskiene, L.; Zubriene, A.; Kairys, V.; Ivanova, J.; Trapencieris, P.; Matulis, D. Saccharin sulfonamides as inhibitors of carbonic anhydrases I, II, VII, XII, and XIII. Biomed. Res. Int. 2014, 2014, No. 638902. (32) Langella, E.; D’Ambrosio, K.; D’Ascenzio, M.; Carradori, S.; Monti, S. M.; Supuran, C. T.; De Simone, G. A combined crystallographic and theoretical study explains the capability of triosephosphate isomerase; TLC, thin-layer chromatography; TMS, tetramethylsilane; TPI, triosephosphate isomerase; WT, wild type ■ Article REFERENCES (1) Ding, Y.; Li, D.; Ding, C.; Wang, P.; Liu, Z.; Wold, E. A.; Ye, N.; Chen, H.; White, M. A.; Shen, Q.; Zhou, J. Regio- and stereospecific synthesis of oridonin D-ring aziridinated analogues for the treatment of triple-negative breast cancer via mediated irreversible covalent warheads. J. Med. Chem. 2018, 61, 2737−2752. (2) Vita, E. D. 10 years into the resurgence of covalent. Future Med. Chem. 2021, 13, 193−210. (3) Ghosh, A. K.; Samanta, I.; Mondal, A.; Liu, W. R. Covalent Inhibition in Drug Discovery. ChemMedChem 2019, 14, 889−906. (4) Shokhen, M.; Hirsch, M.; Khazanov, N.; Ozeri, R.; Perlman, N.; Traube, T.; Vijayakumar, S.; Albeck, A. From catalytic mechanism to rational design of reversible covalent inhibitors of serine and cysteine hydrolases. Isr. J. Chem. 2014, 54, 1137−1151. (5) Singh, J.; Petter, R. C.; Baillie, T. A.; Whitty, A. The resurgence of covalent drugs. Nat. Rev. Drug Discovery 2011, 10, 307−317. (6) McAulay, K.; Hoyt, E. A.; Thomas, M.; Schimpl, M.; Bodnarchuk, M. S.; Lewis, H. J.; Barratt, D.; Bhavsar, D.; Robinson, D. M.; Deery, M. J.; Ogg, D. J.; Bernardes, G. J. L.; Ward, R. A.; Waring, M. J.; Kettle, J. G. Alkynyl benzoxazines and dihydroquinazolines as cysteine targeting covalent warheads and their application in identification of selective irreversible kinase inhibitors. J. Am. Chem. Soc. 2020, 142, 10358−10372. (7) Resnick, E.; Bradley, A.; Gan, J.; Douangamath, A.; Krojer, T.; Sethi, R.; Geurink, P. P.; Aimon, A.; Amitai, G.; Bellini, D.; Bennett, J.; Fairhead, M.; Fedorov, O.; Gabizon, R.; Gan, J.; Guo, J.; Plotnikov, A.; Reznik, N.; Ruda, G. F.; Diaz-Saez, L.; Straub, V. M.; Szommer, T.; Velupillai, S.; Zaidman, D.; Zhang, Y.; Coker, A. R.; Dowson, C. G.; Barr, H. M.; Wang, C.; Huber, K. V. M.; Brennan, P. E.; Ovaa, H.; von Delft, F.; London, N. Rapid covalent-probe discovery by electrophile-fragment screening. J. Am. Chem. Soc. 2019, 141, 8951−8968. (8) Keeley, A.; Petri, L.; Abranyi-Balogh, P.; Keseru, G. M. Covalent fragment libraries in drug discovery. Drug. Discovery Today 2020, 25, 983−996. (9) Gupta, V.; Yang, J.; Liebler, D. C.; Carroll, K. S. Diverse redoxome reactivity profiles of carbon nucleophiles. J. Am. Chem. Soc. 2017, 139, 5588−5595. (10) Sato, M.; Fuchida, H.; Shindo, N.; Kuwata, K.; Tokunaga, K.; Xiao-Lin, G.; Inamori, R.; Hosokawa, K.; Watari, K.; Shibata, T.; Matsunaga, N.; Koyanagi, S.; Ohdo, S.; Ono, M.; Ojida, A. Selective covalent targeting of mutated EGFR(T790M) with chlorofluoroacetamide-pyrimidines. ACS Med. Chem. Lett. 2020, 11, 1137−1144. (11) Martín-Gago, P.; Olsen, C. A. Arylfluorosulfate-based electrophiles for covalent protein labeling: a new addition to the arsenal. Angew. Chem., Int. Ed. 2019, 58, 957−966. (12) Á brányi-Balogh, P.; Petri, L.; Imre, T.; Szijj, P.; Scarpino, A.; Hrast, M.; Mitrovic, A.; Fonovic, U. P.; Nemeth, K.; Barreteau, H.; Roper, D. I.; Horvati, K.; Ferenczy, G. G.; Kos, J.; Ilas, J.; Gobec, S.; Keseru, G. M. A road map for prioritizing warheads for cysteine targeting covalent inhibitors. Eur. J. Med. Chem. 2018, 160, 94−107. (13) Liu, Q.; Sabnis, Y.; Zhao, Z.; Zhang, T.; Buhrlage, S. J.; Jones, L. H.; Gray, N. S. Developing irreversible inhibitors of the protein kinase cysteinome. Chem. Biol. 2013, 20, 146−159. (14) Miseta, A.; Peter, C. Relationship between the occurrence of cysteine in proteins and the complexity of organisms. Mol. Biol. Evol. 2000, 17, 1232−1239. (15) Lagoutte, R.; Remi, P.; Nicolas, W. Covalent inhibitors: an opportunity for rational target selectivity. Curr. Opin. Chem. Biol. 2017, 39, 54−63. (16) Flanagan, M. E.; Abramite, J. A.; Anderson, D. P.; Aulabaugh, A.; Dahal, U. P.; Gilbert, A. M.; Li, C.; Montgomery, J.; Oppenheimer, S. R.; Ryder, T.; Schuff, B. P.; Uccello, D. P.; Walker, G. S.; Wu, Y.; Brown, M. F.; Chen, J. M.; Hayward, M. M.; 9141 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc carboxylic acids to adopt multiple binding modes in the active site of carbonic anhydrases. Chem.Eur. J. 2016, 22, 97−100. (33) Guglielmi, P.; Rotondi, G.; Secci, D.; Angeli, A.; Chimenti, P.; Nocentini, A.; Bonardi, A.; Gratteri, P.; Carradori, S.; Supuran, C. T. Novel insights on saccharin- and acesulfame-based carbonic anhydrase inhibitors: design, synthesis, modelling investigations and biological activity evaluation. J. Enzyme Inhib. Med. Chem. 2020, 35, 1891−1905. (34) Groutas, W. C.; Houser-Archield, N.; Chong, L. S.; Venkataraman, R.; Epp, J. B.; Huang, H.; McClenahan, J. J. Efficient inhibition of human leukocyte elastase and cathepsin G by saccharin. J. Med. Chem. 1993, 36, 3178−3181. (35) Scorza, M. C.; Llado-Pelfort, L.; Oller, S.; Cortes, R.; Puigdemont, D.; Portella, M. J.; Perez-Egea, R.; Alvarez, E.; Celada, P.; Perez, V.; Artigas, F. Preclinical and clinical characterization of the selective 5-HT(1A) receptor antagonist DU-125530 for antidepressant treatment. Br. J. Pharmacol. 2012, 167, 1021−1034. (36) Mauler, F.; Horvath, E. Neuroprotective efficacy of repinotan HCl, a 5-HT1A receptor agonist, in animal models of stroke and traumatic brain injury. J. Cereb. Blood Flow Metab. 2005, 25, 451−459. (37) Goel, P.; Jumpertz, T.; Mikles, D. C.; Ticha, A.; Nguyen, M. T. N.; Verhelst, S.; Hubalek, M.; Johnson, D. C.; Bachovchin, D. A.; Ogorek, I.; Pietrzik, C. U.; Strisovsky, K.; Schmidt, B.; Weggen, S. Discovery and biological evaluation of potent and selective Nmethylene saccharin-derived inhibitors for rhomboid intramembrane proteases. Biochemistry 2017, 56, 6713−6725. (38) Tichá, A.; Collis, B.; Strisovsky, K. The rhomboid superfamily: structural mechanisms and chemical biology opportunities. Trends Biochem. Sci. 2018, 43, 726−739. (39) Alterio, V.; Tanc, M.; Ivanova, J.; Zalubovskis, R.; Vozny, I.; Monti, S. M.; Di Fiore, A.; De Simone, G.; Supuran, C. T. X-ray crystallographic and kinetic investigations of 6-sulfamoyl-saccharin as a carbonic anhydrase inhibitor. Org. Biomol. Chem. 2015, 13, 4064− 4069. (40) Bua, S.; Lomelino, C.; Murray, A. B.; Osman, S. M.; ALOthman, Z. A.; Bozdag, M.; Abdel-Aziz, H. A.; Eldehna, W. M.; McKenna, R.; Nocentini, A.; Supuran, C. T. “A sweet combination”: developing saccharin and acesulfame K structures for selectively targeting the tumor-associated carbonic anhydrases IX and XII. J. Med. Chem. 2020, 63, 321−333. (41) Ivanova, J.; Carta, F.; Vullo, D.; Leitans, J.; Kazaks, A.; Tars, K.; Zalubovskis, R.; Supuran, C. T. N-Substituted and ring opened saccharin derivatives selectively inhibit transmembrane, tumorassociated carbonic anhydrases IX and XII. Bioorg. Med. Chem. 2017, 25, 3583−3589. (42) Hou, R.; He, Y.; Yan, G.; Hou, S.; Xie, Z.; Liao, C. Zinc enzymes in medicinal chemistry. Eur. J. Med. Chem. 2021, 226, No. 113877. (43) Lebraud, H.; Coxon, C. R.; Archard, V. S.; Bawn, C. M.; Carbain, B.; Matheson, C. J.; Turner, D. M.; Cano, C.; Griffin, R. J.; Hardcastle, I. R.; Baisch, U.; Harrington, R. W.; Golding, B. T. Model system for irreversible inhibition of Nek2: thiol addition to ethynylpurines and related substituted heterocycles. Org. Biomol. Chem. 2014, 12, 141−148. (44) Hunter, R. W.; Hughey, C. C.; Lantier, L.; Sundelin, E. I.; Peggie, M.; Zeqiraj, E.; Sicheri, F.; Jessen, N.; Wasserman, D. H.; Sakamoto, K. Metformin reduces liver glucose production by inhibition of fructose-1-6-bisphosphatase. Nat. Med. 2018, 24, 1395−1406. (45) Wright, S. W.; Carlo, A. A.; Carty, M. D.; Danley, D. E.; Hageman, D. L.; Karam, G. A.; Levy, C. B.; Mansour, M. N.; Mathiowetz, A. M.; McClure, L. D.; Nestor, N. B.; McPherson, R. K.; Pandit, J.; Pustilnik, L. R.; Schulte, G. K.; Soeller, W. C.; Treadway, J. L.; Wang, I.-K.; Bauer, P. H. Anilinoquinazoline inhibitors of fructose bisphosphatase bind at a novel allosteric site: synthesis, in vitro characterization, and X-ray crystallography. J. Med. Chem. 2002, 45, 3865−3877. (46) Liu, G. M.; Zhang, Y. M. Targeting FBPase is an emerging novel approach for cancer therapy. Cancer Cell Int. 2018, 18, No. 36. Article (47) Rines, A. K.; Sharabi, K.; Tavares, C. D. J.; Puigserver, P. Targeting hepatic glucose metabolism in the treatment of type 2 diabetes. Nat. Rev. Drug Discovery 2016, 15, 786−804. (48) Huang, Y.; Xu, Y.; Song, R.; Ni, S.; Liu, J.; Xu, Y.; Ren, Y.; Rao, L.; Wang, Y.; Wei, L.; Feng, L.; Su, C.; Peng, C.; Li, J.; Wan, J. Identification of the new covalent allosteric binding site of fructose1,6-bisphosphatase with disulfiram derivatives toward glucose reduction. J. Med. Chem. 2020, 63, 6238−6247. (49) Huang, Y.; Wei, L.; Han, X.; Chen, H.; Ren, Y.; Xu, Y.; Song, R.; Rao, L.; Su, C.; Peng, C.; Feng, L.; Wan, J. Discovery of novel allosteric site and covalent inhibitors of FBPase with potent hypoglycemic effects. Eur. J. Med. Chem. 2019, 184, No. 111749. (50) Ma, L.; Zhang, H.; Humphreys, W. G.; Zhu, M. Detection and structural characterization of glutathione-trapped reactive metabolites using liquid chromatography-high-resolution mass spectrometry and mass defect filtering. Anal. Chem. 2007, 79, 8333−8341. (51) Frisch, M. J. T.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford CT, 2013. (52) Hong, J. R.; Choi, Y. J.; Keum, G.; Nam, G. Synthesis and diabetic neuropathic pain-alleviating effects of 2N-(pyrazol-3-yl)methylbenzo[d]isothiazole-1,1-dioxide derivatives. Bioorg. Med. Chem. 2017, 25, 4677−4685. (53) (a) Romani, S.; Bovermann, G.; Moroder, L.; Wunsch, E. N-(2Nitrobenzenesulphenyl)-saccharin: A Useful Reagent in Peptide Synthesis. Synthesis 1985, 1985, 512−513. (b) Hunter, R. W.; Hughey, C. C.; Lantier, L.; Sundelin, E. I.; Peggie, M.; Zeqiraj, E.; Sicheri, F.; Jessen, N.; Wasserman, D. H.; Sakamoto, K. Metformin reduces liver glucose production by inhibition of fructose-1-6bisphosphatase. Nat. Med. 2018, 24, 1395−1406. (54) Tokunaga, K.; Sato, M.; Kuwata, K.; Miura, C.; Fuchida, H.; Matsunaga, N.; Koyanagi, S.; Ohdo, S.; Shindo, N.; Ojida, A. Bicyclobutane carboxylic amide as a cysteine-directed strained electrophile for selective targeting of proteins. J. Am. Chem. Soc. 2020, 142, 18522−18531. (55) Jöst, C.; Nitsche, C.; Scholz, T.; Roux, L.; Klein, C. D. Promiscuity and selectivity in covalent enzyme inhibition: a systematic study of electrophilic fragments. J. Med. Chem. 2014, 57, 7590−7599. (56) Lonsdale, R.; Burgess, J.; Colclough, N.; Davies, N. L.; Lenz, E. M.; Orton, A. L.; Ward, R. A. Expanding the armory: predicting and tuning covalent warhead reactivity. J. Chem. Inf. Model. 2017, 57, 3124−3137. (57) Mancuso, A. J.; Swern, D. Activated dimethyl sulfoxide: useful reagents for synthesis. Synthesis 1981, 1981, 165−185. (58) Veverka, K. A.; Johnson, K. L.; Dennis, C. M.; James, J. L.; Stephen, N. Inhibition of aldehyde dehydrogenase by disulfiram and its metabolite methyl diethylthiocarbamoyl-sulfoxide. Biochem. Pharmacol. 1997, 53, 511−518. (59) Xu, Y.-x.; Huang, Y.; Song, R.; Ren, Y.; Chen, X.; Zhang, C.; Mao, F.; Li, X.; Zhu, J.; Ni, S.; Wan, J.; Li, J. Development of disulfide-derived fructose-1, 6-bisphosphatase (FBPase) covalent inhibitors for the treatment of type 2 diabetes. Eur. J. Med. Chem. 2020, 203, No. 112500. (60) Solca, F.; Dahl, G.; Zoephel, A.; Bader, G.; Sanderson, M.; Klein, C.; Kraemer, O.; Himmelsbach, F.; Haaksma, E.; Adolf, G. R. 9142 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143 Journal of Medicinal Chemistry pubs.acs.org/jmc Target binding properties and cellular activity of afatinib (BIBW 2992), an irreversible ErbB family blocker. J. Pharmacol. Exp. Ther. 2012, 343, 342−350. (61) Sinnokrot, M. O.; Valeev, E. F.; Sherrill, C. D. Estimates of the ab initio limit for pi−pi interactions: The benzene dimer. J. Am. Chem. Soc. 2002, 124, 10887−10893. (62) Shi, Z.; Olson, C. A.; Kallenbach, N. R. Cationπ interaction in model α-helical peptides. J. Am. Chem. Soc. 2002, 124, 3284−3291. (63) Carnaroglio, D.; Martina, K.; Palmisano, G.; Penoni, A.; Domini, C.; Cravotto, G. One-pot sequential synthesis of isocyanates and urea derivatives via a microwave-assisted Staudinger-aza-Wittig reaction. Beilstein J. Org. Chem. 2013, 9, 2378−2386. (64) Flanagan, M. E.; Abramite, J. A.; Anderson, D. P.; Aulabaugh, A.; Dahal, U. P.; Gilbert, A. M.; Li, C.; Montgomery, J.; Oppenheimer, S. R.; Ryder, T.; Schuff, B. P.; Uccello, D. P.; Walker, G. S.; Wu, Y.; Brown, M. F.; Chen, J. M.; Hayward, M. M.; Noe, M. C.; Obach, R. S.; Philippe, L.; Shanmugasundaram, V.; Shapiro, M. J.; Starr, J.; Stroh, J.; Che, Y. Chemical and computational methods for the characterization of covalent reactive groups for the prospective design of irreversible inhibitors. J. Med. Chem. 2014, 57, 10072−10079. (65) Chen, D.; Guo, D.; Yan, Z.; Zhao, Y. Allenamide as a bioisostere of acrylamide in the design and synthesis of targeted covalent inhibitors. MedChemComm 2018, 9, 244−253. (66) Han, X.; Huang, Y.; Zhang, R.; Xiao, S.; Zhu, S.; Qin, N.; Hong, Z.; Wei, L.; Feng, J.; Ren, Y.; Feng, L.; Wan, J. New insight into the binding modes of TNP-AMP to human liver fructose-1,6-bisphosphatase. Spectrochim. Acta, Part A 2016, 165, 155−160. (67) Huang, Y.; Xu, Y.; Song, R.; Ni, S.; Liu, J.; Xu, Y.; Ren, Y.; Rao, L.; Wang, Y.; Wei, L.; Feng, L.; Su, C.; Peng, C.; Li, J.; Wan, J. Identification of the new covalent allosteric binding site of fructose1,6-bisphosphatase with disulfiram derivatives toward glucose reduction. J. Med. Chem. 2020, 63, 6238−6247. (68) Park, J. B.; Park, H.; Son, J.; Ha, S. J.; Cho, H. S. Structural study of monomethyl fumarate-bound human GAPDH. Mol. Cells 2019, 42, 597−603. (69) Kornberg, M. D.; Bhargava, P.; Kim, P. M.; Putluri, V.; Snowman, A. M.; Putluri, N.; Calabresi, P. A.; Snyder, S. H. Dimethyl fumarate targets GAPDH and aerobic glycolysis to modulate immunity. Science 2018, 360, 449−453. (70) Blonski, C.; De Moissac, D.; Perie, J.; Sygusch, J. Inhibition of rabbit muscle aldolase by phosphorylated aromatic compounds. Biochem. J. 1997, 323, 71−77. (71) Kashem, M. A.; Nelson, R. M.; Yingling, J. D.; Pullen, S. S.; Prokopowicz, A. S.; Jones, J. W.; Wolak, J. P.; Rogers, G. R.; Morelock, M. M.; Snow, R. J.; Homon, C. A.; Jakes, S. Three mechanistically distinct kinase assays compared: Measurement of intrinsic ATPase activity identified the most comprehensive set of ITK inhibitors. SLAS Discovery 2007, 12, 70−83. (72) Kabsch, W. Xds. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2010, 66, 125−132. (73) McCoy, A. J.; Grosse-Kunstleve, R. W.; Adams, P. D.; Winn, M. D.; Storoni, L. C.; Read, R. J. Phaser crystallographic software. J. Appl. Crystallogr. 2007, 40, 658−674. (74) Emsley, P.; Lohkamp, B.; Scott, W. G.; Cowtan, K. Features and development of Coot. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2010, 66, 486−501. (75) Adams, P. D.; Afonine, P. V.; Bunkoczi, G.; Chen, V. B.; Davis, I. W.; Echols, N.; Headd, J. J.; Hung, L. W.; Kapral, G. J.; GrosseKunstleve, R. W.; McCoy, A. J.; Moriarty, N. W.; Oeffner, R.; Read, R. J.; Richardson, D. C.; Richardson, J. S.; Terwilliger, T. C.; Zwart, P. H. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2010, 66, 213−221. (76) Moriarty, N. W.; Grosse-Kunstleve, R. W.; Adams, P. D. Electronic ligand builder and optimization workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2009, 65, 1074−1080. Article (77) Chen, V. B.; Arendall, W. B., 3rd; Headd, J. J.; Keedy, D. A.; Immormino, R. M.; Kapral, G. J.; Murray, L. W.; Richardson, J. S.; Richardson, D. C. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2010, 66, 12−21. Recommended by ACS Preferred Conformation-Guided Discovery of Potent and Orally Active HIF Prolyl Hydroxylase 2 Inhibitors for the Treatment of Anemia Yue Wu, Xiaojin Zhang, et al. JUNE 27, 2023 JOURNAL OF MEDICINAL CHEMISTRY READ Cyclic Peptidic Furin Inhibitors Developed by Combinatorial Chemistry Agata Gitlin-Domagalska, Krzysztof Rolka, et al. MARCH 15, 2023 ACS MEDICINAL CHEMISTRY LETTERS READ Anticancer Properties of Hexosamine Analogs Designed to Attenuate Metabolic Flux through the Hexosamine Biosynthetic Pathway Christopher T. Saeui, Kevin J. Yarema, et al. JANUARY 10, 2023 ACS CHEMICAL BIOLOGY READ Discovery of NAFLD-Improving Agents by Promoting the Degradation of Keap1 Mingming Qi, Jianbo Sun, et al. JUNE 30, 2023 JOURNAL OF MEDICINAL CHEMISTRY READ Get More Suggestions > 9143 https://doi.org/10.1021/acs.jmedchem.2c00336 J. Med. Chem. 2022, 65, 9126−9143